Next Article in Journal
Three-Year Analysis of Adjuvant Therapy in Postoperative Melanoma including Acral and Mucosal Subtypes
Previous Article in Journal
FUCA1: An Underexplored p53 Target Gene Linking Glycosylation and Cancer Progression
Previous Article in Special Issue
A New Hypothesis Describing the Pathogenesis of Oral Mucosal Injury Associated with the Mammalian Target of Rapamycin (mTOR) Inhibitors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Non-Receptor Tyrosine Kinases: Their Structure and Mechanistic Role in Tumor Progression and Resistance

by
Abdulaziz M. Eshaq
1,2,
Thomas W. Flanagan
3,
Sofie-Yasmin Hassan
4,
Sara A. Al Asheikh
2,
Waleed A. Al-Amoudi
2,
Simeon Santourlidis
5,
Sarah-Lilly Hassan
6,
Maryam O. Alamodi
2,
Marcelo L. Bendhack
7,
Mohammed O. Alamodi
2,
Youssef Haikel
8,9,10,
Mossad Megahed
11 and
Mohamed Hassan
8,9,12,*
1
Department of Epidemiology and Biostatistics, Milken Institute School of Public Health, George Washington University, Washington, DC 20052, USA
2
College of Medicine, Alfaisal University, Riyadh 11533, Saudi Arabia
3
Department of Pharmacology and Experimental Therapeutics, LSU Health Sciences Center, New Orleans, LA 70112, USA
4
Department of Pharmacy, Faculty of Science, Heinrich-Heine University Duesseldorf, 40225 Duesseldorf, Germany
5
Institute of Cell Therapeutics and Diagnostics, University Medical Center of Duesseldorf, 40225 Duesseldorf, Germany
6
Department of Chemistry, Faculty of Science, Heinrich-Heine University Duesseldorf, 40225 Duesseldorf, Germany
7
Department of Urology, Red Cross University Hospital, Positivo University, Rua Mauá 1111, Curitiba 80030-200, Brazil
8
Institut National de la Santé et de la Recherche Médicale, University of Strasbourg, 67000 Strasbourg, France
9
Department of Operative Dentistry and Endodontics, Dental Faculty, 67000 Strasbourg, France
10
Pôle de Médecine et Chirurgie Bucco-Dentaire, Hôpital Civil, Hôpitaux Universitaire de Strasbourg, 67000 Strasbourg, France
11
Clinic of Dermatology, University Hospital of Aachen, 52074 Aachen, Germany
12
Research Laboratory of Surgery-Oncology, Department of Surgery, Tulane University School of Medicine, New Orleans, LA 70112, USA
*
Author to whom correspondence should be addressed.
Cancers 2024, 16(15), 2754; https://doi.org/10.3390/cancers16152754
Submission received: 29 May 2024 / Revised: 30 June 2024 / Accepted: 29 July 2024 / Published: 2 August 2024
(This article belongs to the Special Issue Tumor Microenvironment and Treatment Resistance)

Abstract

:

Simple Summary

Protein tyrosine kinases (PTKs) are classified into two groups: one group includes tyrosine kinases, and the second group includes serine/threonine kinases. The group of tyrosine kinases includes both receptor tyrosine kinases (RTKs) and non-receptor tyrosine kinases (NRTKs) that function as “on” or “off” switches for many cellular functions. NRTKs are kinase enzymes which are overexpressed and activated in many cancer types and regulate variable cellular functions, including cell growth and progression and their dependent mechanisms and the associated signaling pathways. Thus, targeting NRTKs is of great interest to improve the treatment strategy of different tumor types.

Abstract

Protein tyrosine kinases (PTKs) function as key molecules in the signaling pathways in addition to their impact as a therapeutic target for the treatment of many human diseases, including cancer. PTKs are characterized by their ability to phosphorylate serine, threonine, or tyrosine residues and can thereby rapidly and reversibly alter the function of their protein substrates in the form of significant changes in protein confirmation and affinity for their interaction with protein partners to drive cellular functions under normal and pathological conditions. PTKs are classified into two groups: one of which represents tyrosine kinases, while the other one includes the members of the serine/threonine kinases. The group of tyrosine kinases is subdivided into subgroups: one of them includes the member of receptor tyrosine kinases (RTKs), while the other subgroup includes the member of non-receptor tyrosine kinases (NRTKs). Both these kinase groups function as an “on” or "off" switch in many cellular functions. NRTKs are enzymes which are overexpressed and activated in many cancer types and regulate variable cellular functions in response to extracellular signaling-dependent mechanisms. NRTK-mediated different cellular functions are regulated by kinase-dependent and kinase-independent mechanisms either in the cytoplasm or in the nucleus. Thus, targeting NRTKs is of great interest to improve the treatment strategy of different tumor types. This review deals with the structure and mechanistic role of NRTKs in tumor progression and resistance and their importance as therapeutic targets in tumor therapy.
Keywords:
PTK; NRTKs; Ack; Jak; Fes; Fak; Tec; Src; Csk; Abl; Syk kinases

1. Introduction

The expression and activation of protein tyrosine kinases (RTKs) are common in most eukaryotic organisms. These PTKs are involved in the regulation of different cellular processes under normal and pathological conditions [1,2,3]. However, aberrant activation of these PTKs in cancer is mostly associated with disease progression and drug resistance [3,4]. Therefore, PTKs are considered not only as important signaling molecules but also as important drug targets in many diseases, especially cancer [1,2]. The family members of the PTK family are characterized by their diversity in structure and function; in addition, they play an essential role in the regulation of different cellular functions [2,3]. The catalytic kinase domains of PTKs are characterized by the ability to phosphorylate substrate proteins at tyrosine residues. The group of tyrosine kinases is subdivided into subgroups: one of them includes the member of receptor tyrosine kinases (RTKs), while the other subgroup includes the member of non-receptor tyrosine kinases (NRTKs) [4,5]. The main function of RTKs is to rapidly and reversibly phosphorylate protein substrates on their serine, threonine, or tyrosine residues, leading to a significant alteration in protein characteristics, including conformation and interaction to drive different cellular processes [3,4]. The two groups of PTKs include the group of tyrosine kinases and the serine/threonine kinases [6]. Receptor serine/threonine kinases (RSTKs) are extracellular transmembrane proteins with cytoplasmic kinase domains. The extracellular domain serves as a ligand-binding domain, while the cytoplasmic domain serves as a bridge between extracellular signaling receptors for members of transforming growth factor-β to maintain the progression of signal transduction processes [7,8]. Tyrosine kinases are a large family of kinases that include both RTKs and NRTKs that function as “on” or “off” switches in many cellular functions [1,9]. Although both RTKs and NRTKs are similar in function, their structures are different [10]. In contrast to RTKs, the NRTKs lack both transmembrane and extracellular domains [11,12], while RTKs are cell surface receptors that functionally transmit growth factors, cytokines, and hormone signaling to the cytoplasm and nucleus [13,14]. NRTKs are both cytoplasmic and nuclear proteins; the localization of the cytoplasmic NRTK proteins is either free or anchored to the inner side of the cell membrane [15,16]. The main function of the cytoplasmic NRTK protein is to mediate intracellular signals resulting from extracellular receptor-dependent activation [15,17].
NRTKs are enzymes with kinase activity whose activation is closely associated with significant changes in annotated molecular and cellular functions [15,16]. Based on their sequence similarity, particularly withing their kinase domains, the NRTKs are grouped into different subfamilies, which are implicated in the regulation of different cellular functions, including cell growth, cell division, and immune regulation [15,18]. In this review, the structure and the mechanistic role of NRTKs in tumor progression and drug resistance are discussed.

2. Structure of Non-Receptor Tyrosine Kinases

In contrast to the RTKs, the members of NRTKs lack both the extracellular ligand-binding domain and the transmembrane-spanning region, which are characteristic of RTKs. The different members of NRTKs are either localized in the cytoplasm or anchored to the cell membrane via significant modification in their amino terminus [15,18].
The most common NRTKs include Ack, Jak, Fes, Fak, Tec, Src, Csk, Abl, and Syk kinases [15,18]. Apart from their classification, NRTKs share distinct sequence similarity within their kinase domains [15,18,19]. These kinase domains are highly conserved, and their catalytic domains are like that of Ser/Thr protein kinases [15,18,19].
Of note, Ser/Thr protein kinases are characterized by their lobes on their N- and C-terminal domains. The lobe of the N-terminal domain serves as a binding site for the phosphate groups of adenosine triphosphate (ATP), and the lobe of the C-terminal domain serves as a substrate-binding site for both ATP and peptides [20,21]. The activation loop of the C-terminal domain of NRTKs contains either Tyr, Ser, or Thr residues. The phosphorylation of most NRTKs is mediated by either autophosphorylation or other members of NRTKs at tyrosine sites in the activation loop, and the level of their phosphorylation is mostly associated with significant activity [22,23].
Beside the catalytic domains, NRTKs also contain non-catalytic domains. The main function of non-catalytic domains is to regulate the activation of NRTKs via a mechanism mediated by both inter- and intramolecular interactions of PTKs [15,23,24]. The most common non-catalytic domains of NRTKs are the Src homology 2 (SH2) and Src homology 3 (SH3) domains [15,23,24]. The SH2 domain is characterized by its affinity for binding to proline-rich motifs existing in many proteins [15,25,26]. However, the classification of NRTKs into distinct families relies on the molecular analysis of the domain structure based on the variation in amino acid sequences and the genomic organization of the kinase domains [19]. The structure of the most common NRTKs families is outlined in Figure 1.

2.1. Ack Family

The Ack is a large protein of 120 kDa whose kinase activity can be mediated by the phosphorylation of its tyrosine residues. The Ack family is comprised of two members, including Ack1, the activated Cdc42-associated kinase 1, and Tnk1, the thirty-eight-negative kinase-1 [27,28,29]. Ack1 is characterized by its unique structure, which differs significantly from that of the other members of NRTKs. In contrast to homologous proteins, the human Ack1/Tnk2 is characterized by its domain architecture that includes an N-terminal sterile alpha motif (SAM) domain [19,30,31,32], an SH3 domain, and a Cdc42/Rac-interactive domain (CRIB), as well as several functionally relevant regions located in the C-terminal region of Ack1 [19,30,33]. These functional regions include multiple proline-rich sequences, a clathrin-binding motif (CB) [34], a region that shares a high homology with Mig6 [19,35], and a ubiquitin binding domain (UBS) [36]. Besides being the only NRTK with the SH3 domain located to the C-terminal of the kinase, the Ack family kinases have the unique CRIB domain [19,30,31]. The catalytic kinase domain of Ack is in the C-terminal region and is closely followed by the SH3 domain, while the SH3 domain is directly followed by the CRIB domain and is characterized by its GTP-dependent binding to Cdc42, but not binding to those of Rho family guanosine triphosphatase (GTPase) [19,30,37]. The homology of both Ack and Tnk ends directly after the SH3 domain (Tnk lacks the CRIB domain that is characteristic of Ack1 [28]). Ack1 contains an arrestin-like clathrin-binding region that is located directly after the CRIB domain [22,38]. However, the multidomain structure of the Ack1 family not only facilitates its localization to different cellular compartments but also promotes its association with different proteins that enhance its functional diversity [22,39].
Although the conformation of the activation loop of both the phosphorylated and non-phosphorylated forms of Ack is similar, the similarity of the Ack forms is that they cannot cover the substrate binding site, suggesting that phosphorylation of the activation loop does not appear to play a stimulatory role in the activation of Ack kinase [40,41].
The functional analysis of the Ack family members provided insight into the mechanisms regulating the activation of Ack both in normal and cancer cells [27]. The identification of 10 amino acid sequences rich in proline residues within Ack1 up-stream of the kinase recognition segment of the Mig6 homology region (MHR) has been shown to play an essential in the interaction between the proline-rich sequence and SH3 domain to stabilize the auto-inhibited state of the Ack1 kinase via a mechanism mediated by the MHR-dependent inhibition of the C-terminal region of the kinase domain [22,30,31]. Thus, Ack1 activation is a consequence of the interaction of RTKs with the MHR of Ack1 [22,31,42]. The complete activation of Ack1 has been reported to be attributed to the amino-terminal SAM domain-dependent mechanism, including membrane localization and the symmetric dimerization of Ack1, leading to Ack1 transphosphorylation and the activation of Ack1 [22,27]. To that end, Ack1 is expected to switch to different modes of kinase activation and to be adapted to cellular requirements.
The identification of the substrates of Ack1 has taken place both in normal and cancer cells. For example, the phosphorylation of Wiskott–Aldrich syndrome protein (WASP) by Ack1 has been reported to promote its actin remodeling activity [28], while the phosphorylation of p130Cas, a component of focal adhesion and a member of the Cas (Crk-associated substrate) family, by Ack1 is associated with tumor invasion, promoting cell spreading [43] and cell migration [44]. In addition to its role in the regulation of cell adhesion and migration, Ack1 has also been reported to phosphorylate the androgen receptor (AR), leading to the enhancement of AR-mediated gene transcription [45,46]. The domain architecture of the human Ack and its homologous proteins in different organisms are shown in Figure 2.
Ack1 resides in the cellular cytoplasm downstream of a broad range of receptors such as integrins, muscarinic M3 receptors, platelet-derived growth factor receptors [PDGFR], Axl receptor tyrosine kinases, insulin receptors (IRs), Mer receptor tyrosine kinases, Trk, and EGFR [34,47,48]. The activation of Ack1 results from the interaction of ligand-activated RTKs with the MHR to overcome the autoinhibitory interaction of the MHR with its own kinase domain [19,22,35]. However, the complete activation of Ack1 is mediated by the SAM domain-dependent mechanism to facilitate membrane localization and the symmetric dimerization of Ack1, leading to its transphosphorylation and activation [22,35].
Ack1 kinase as an oncogenic protein relies on the ability of activated Ack1 to promote tumor growth in vivo [49]. The oncogenicity of Ack1 is attributed to the aberrant activation of Ack1 by different mechanisms, including gene amplification, missense mutation of Ack1, and ligand-dependent activation of RTKs [27,50,51]. The gene amplification of Ack1 has been reported in different tumor types, including cervical, ovarian, lung squamous, head and neck squamous cell carcinomas, breast, prostate, and gastric cancers [27,50,51]. Wide genome analysis revealed the existence of both missense and nonsense mutations within the domains of the Ack1 gene [22,52].
The functional analysis of Ack1/Tnk2 helped address the mechanisms whereby these kinases regulate tumor growth and resistance to anticancer agents. The main function of activated Ack1 is to transduce the extracellular signals from the stimulated RTKs by their corresponding ligands [27,53]. RTKs are G-protein coupled receptors (GPCRs), such as platelet-derived growth factor receptor (PDGFR), insulin receptor (IR), and epidermal growth factor receptor (EGFR) [53,54,55,56]. In cancer cells, activated Ack1 mediates the phosphorylation of Wwox protein that is thought to play an essential role in triggering the activation of the Akt pathway and androgen receptor (AR), leading to tumor growth [45,46]. Also, the genetic alteration of Ack1 through gene amplification and somatic mutation can cause the neoplastic transformation of a variety of human malignancies [57]. For example, Somatic mutation in the ubiquitin-binding domains (UBDs) of Ack1 has been shown to mediate EGFR regulation in renal carcinoma cells [54,55,56]. The development of tumor resistance by Ack1 is attributed to its function as an epigenetic regulator [27,58,59,60]. Ack1-induced tumor resistance to anticancer agents is mediated through the interaction of Ack1 with the estrogen receptor (ER)/histone demethylase KDM3A (JHDM2a) complex. Consequently, this complex is responsible for the modification of KDM3A via tyrosine phosphorylation, which leads to the regulation of the transcriptional outcome at homeobox A1 (HOXA1) [27,58]. The epigenetic reprogramming of cell cycle genes by Ack1 has been reported to promote breast cancer resistance to CDK4/6 inhibitors [59,61] and the inhibitor of the estrogen receptor modulator, tamoxifen [62,63]. Loss of Ack1 has been shown to be associated with the upregulation of EGFR and to enhance tumor resistance to BRAF inhibitors [64].
The inhibition of the Ack1-mediated phosphorylation of the C-terminal Src kinase has been found to counteract prostate cancer resistance to immune checkpoint blockade [60] and to overcome the acquired resistance of EGFR mutant non-small-cell lung cancer (NSCLC) cells to Osimertinib, the inhibitor of EGFR [64]. Importantly, the combination of ACK1 and KIT inhibitors was found to have a positive therapeutic impact on imatinib-resistant gastrointestinal stromal tumors via anti-proliferative and -migration effect-dependent mechanisms [65], as well as being superior to KRAS-mutant-based therapies in NSCLC [66].
In addition to its role in tumor progression and treatment resistance, the involvement of Ack1 in tumor recurrence has been reported in human hepatocellular carcinoma (HCC) [67] and in prostate cancer [68,69].

2.2. Jak/Janus Family

The Jak/Janus family includes four members (JAK1, 2, 3 and TYK2) that are quite different from other NRTKs via their two kinase domains in their C-terminal half [11,15,70]. These two kinase domains are regulated so that one is functional, whereas the other serves as pseudo-kinase domain [70,71]. In contrast to the C-terminal region, the N-terminal region of the Jak family contains stretches of highly conserved sequences that are known as JH domains [71,72,73]. These sequences represent the four-point-one, ezrin, radixin, moesin (FERM), and SH2 domains of Jak family members [74,75,76]. The structures of Jaks include Jak1, Jak2, and Jak3 and TYK2, as outlined in Figure 1.
Unlike other receptor tyrosine kinases, the cytokine receptors lack intrinsic protein kinase domains; therefore, the cytokine receptor cannot cause rapid tyrosine phosphorylation of intracellular proteins [77,78,79]. Consequently, the cytokine receptors recruit and activate a variety of NRTKs, including the JAK family, to induce downstream signaling pathways [78,80]. Cytokine receptors comprise two or more receptor subunits, including γ chain and gp130, that are associated with a JAK monomer [81,82]. Upon receptor ligation by a cognate ligand, the receptor subunits are either realigned or oligomerized to recruit and associate JAKs to facilitate their transphosphorylation and subsequently corresponding activation [83,84]. Thus, once the recruiting and activation of JAKs have been completed, they become able to phosphorylate tyrosine residues within the cytoplasmic regions of the cytokine receptors and thereby facilitate the interaction of the downstream adaptor and effector proteins containing SH2 domains with the cytoplasmic region of cytokine receptors [85,86].
The Jak family members Jak1, Jak2, and Tyk2 are omnipresent in mammals, while Jak3 expression is restricted predominantly to hematopoietic cells [87,88]. JAKs are located mainly in the cytosol, and their expression is cytokine receptors-independent [87,88,89]. Because of their close association with cytokine receptors, the members of the Jak family are localized along with their associated receptors either in endosomes or in the plasma membrane [81,82,83,84,85,86,87,88,89].
The link between Jak family members and cytokine signaling is widely reported [89,90,91,92]. The ligation of cytokines, namely interleukin (IL)-2, IL-4, IL-7, IL-9, IL-15, and IL-21, to the corresponding receptors on the surface of target cells induces their auto-activation by transphosphorylation, leading to the activation of Jak1-dependent signaling pathways [93,94,95,96]. The enhancement of Jak3 activation by the abovementioned cytokines results from the binding of Jak3 to the γ chain subunit of the cytokine receptors [95,96,97,98]. Jak1 is activated by many factors, including IL-6, IL-11, oncostatin M, leukemia inhibitory factor (LIF), ciliary neurotrophic factor (CNF), granulocyte colony-stimulating factor (G-CSF), and IFNs. The abovementioned factors-induced stimulation of Jak1 is mediated by the subunit gp130 of cytokine receptors [99,100,101,102], while the activation of Jak2 is mediated by the interactions of Jak with its cognate transmembrane cytokine receptor proteins in response to growth hormone (GH), prolactin (PRL), erythropoietin (EPO), thrombopoietin (TPO), and the family of cytokines that signal through the IL-3 receptor (IL-3, IL-5, and granulocyte–macrophage colony-stimulating factor, as well as the GM-CSF-dependent mechanism) [103,104].
Tyk2 is the first member of the Jak family that is involved in the modulation of IFN signaling-dependent mechanisms [97,98]. Although Tyk2 is essential for the modulation of IL-12 signaling-dependent mechanisms, its role in the modulation of IFN-α/β signaling or cytokines, particularly those using gp130, is not essential [104,105,106].
In addition to their significant role leading the regulation of immune cells’ growth and fate [107], the JAK-STAT pathway has been shown to play key roles in the regulation of tumor development and plasticity [108]. The structure and conserved phosphorylation sites of the JAK family are outlined as shown in Figure 3.

2.3. Fes Kinases

Feline sarcoma (Fes) and Fes-related (Fer) kinases are a subgroup of NRTKs with similarity to viral oncogenes. These viral oncogenes are derived from feline sarcoma virus (v-FES) and avian Fujinami poultry sarcoma virus (v-fps). Although the ubiquitous expression of Fer is well established, the expression of FES is only limited to myeloid hematopoietic, neuronal, epithelial, and vascular endothelial cells [15,109,110].
Fes kinases are characterized by their unique amino-terminal Fes/Fer/Cdc42-interacting protein homology (FCH) domain in addition to constituting three coiled-coil motifs, a central SH2 domain, and a kinase domain that is in the C-terminal region. The FCH domain lies adjacent to a coiled-coil domain that is like the BAR domain that forms a functional unit that is known as the FCH-Bin–Amphiphysin–Rvs (F-BAR) domain [111,112].
Although Fes kinase lacks the SH3 domain, the repression of Fes is mediated via the interaction between the SH2 and kinase domain-dependent mechanism [112,113], while the activation of Fes kinase is mediated through the phosphorylation of both Tyr713 and Tyr811 residues [112,113,114]. The structure and the functional domains of the NRTK Fes family members are outlined in Figure 4.

2.4. Fak Family

The Fak family includes Fak that is expressed ubiquitously and other family members (Pyk2, Cak-beta, Cadtk, Raftk, and Fak2), which are expressed mainly in different organs, including the brain, liver, lung, kidney, and hematopoietic cells [114,115]. Fak family members are flanked with extensive N- and C-terminal regions [116,117,118]; the N-terminal region of Fak family members contains the four-point-one, ezrin, radixin, and moesin (FERM) domain that is followed by the kinase domain. FERM is essential for the modulation the interactions of Fak family members with transmembrane receptors [119,120,121]. The FERM domain of Fak can also mediate interactions of Fak with integrins and/or RTKs [122,123]. The C-terminal region of the Fak family possesses a region that is implicated in focal adhesion targeting (FAT) [124,125,126]. The role of paxillin and talin in the modulation of the FAT-dependent targeting of Fak has been reported [127,128,129,130].
The involvement of the Fak family in the regulation of different cellular functions and processes has been reported in several studies [121,131,132,133]. Apart from the catalytic activity of FAK in tumor cells, its subcellular localization of Fak in both the cytoplasm and nucleus can be crucial for the transcriptional regulation of chemokines that can promote the tumor microenvironment (TME) to attenuate the antitumor properties of the host [121,132,133]. Accumulated evidence suggests an important role for Fak family members in malignant processes [134]. Several pathological studies indicate an overexpression of Fak protein in many types, including the colon, ovary, pancreas, and prostate [133]. As expected, the overexpression of Fak was found to be correlated with more aggressive and invasive breast carcinomas [133]. Also, PTK2, the gene that encodes Fak, is highly amplified in human breast cancer and is associated with metastasis-free survival [135,136]. The structure and functional domains of Fak family kinases and Src-FAK-dependent signaling pathways are outlined in Figure 5A and Figure 5B, respectively.

2.5. Tec Family

The Tec family is an NRTK that contains five members with high similarity in their functional domains. The members of Tec family kinases are comprised of the Pleckstrin homology domain (PH) and a Tec homology domain (TH) that is divided into the cysteine-rich, Zn2+-binding Btk motif and a proline-rich motif [137,138,139]. The PH and TH domains are followed by SH3 and SH2 domains that are consequently followed by the kinase domain [140,141,142]. Unlike other Tech family members, the Bmx and Txk kinases do not constitute the typical structure of most of the Tec family members. Both the TH and SH3 domains of Bmx kinase are known as the TH- and SH3-like domains [143,144]. In Txk, both the PH and TH domains are replaced by a unique cysteine string motif containing serine and threonine sites, which is the common palmitoylation [145,146,147].
Although both Bmx and Tec are widely expressed in different tissues and cell types, the expression of other Tec family members is noted mainly in hematopoietic and non-hematopoietic cells [147,148]. Itk is expressed mainly in T and NK cells, and the expression of Btk is common in B lymphocytes and cells of erythromyeloid lineage, and Txk has been shown to be expressed in T helper Type1 cells and to regulate the production of interferon γ in human T lymphocytes [149,150].
The activation of Tec kinases is mediated by membrane targeting via the interaction of their PH domains with phosphatidylinositol (3,4,5)-trisphosphate (PIP3) or with other proteins, as well as in response to the phosphorylation of tyrosine residues within the kinase activation loop [141,151]. The structure and functional domains of Tec kinases are demonstrated in Figure 6.

2.6. Src Family

The Src is one of the largest family members of NRTKs that is divided into two classes. One of these classes includes the tyrosine kinases with a broad expression range (e.g., Fyn, Yes), while the other class includes kinases with a limited expression range (e.g., Fgr, Lyn, Hck, Lck, Blk, Yrc) [15]. Src family proteins of vertebrates have a similar structure with a molecular mass ranging from 52 to 62 kDa [152,153]. This group of kinases comprise six distinct functional domains that consist of an N-terminal sequence constitute SH4 region of 15 to 17 amino acids residues, including signals for fatty acid modification sites, followed by a unique domain that possesses a low degree of intra-family homology next to the N-terminal region, followed by the SH3, SH2, and tyrosine kinase domains [152,153,154]. The C-terminal part of Src protein contains tyrosine residues, which are essential for the regulation of Src activity [155,156,157].
The analysis of the protein sequence of the SRC family of kinases (SFKs) revealed the presence of eight members, including BLK, FGR, FYN, HCK, LCK, LYN, SRC, CHK, and YES. The members of SKFs are further divided into two subfamilies: the SRC-related subfamily that is known as Src A, that includes the FGR, FYN, SRC, and YES kinases, while the other subfamily, namely the LYN-related subfamily, known as Src-B, includes the BLK, HCK, LCK, and LYN kinases [158,159,160].
The C-terminal of the members of the Src A and its homologous kinase Scr-B is characterized by their function as an endogenous inhibitor for Src family protein tyrosine kinases (SFKs) [158,159,160]. The Csk family of NRTKs includes Csk and Chk (also known as Ctk, Ntk, Chk, Hyl, and Lsk). Although Csk is expressed in all organs, it is predominantly found in the thymus and spleen [15,161], while Chk is expressed primarily in the brain and hematopoietic cells [161,162,163]. In addition to the high homology of the amino acid sequences of the Csk family with those of the Src family, the domain structure of the Csk family includes an SH3 domain, followed by an SH2 domain and a tyrosine kinase domain, evidence of its high similarity with the Src family [161,164]. Apart from their higher structural similarity, the two Scr-A and Scr-B kinases underly several divergences in their structures [15]. For example, the Csk family lacks the N-terminal unique domain the constitutes the fatty acid modification sites. This unique domain is essential to anchor the Src family to the membrane [155,164,165]. In addition to the absence of the C-terminal regulatory site, the Csk family possesses no tyrosine residue in its activation loop [24,161,164]. The structural and functional domains of Src kinases are shown Figure 7.

2.7. BRK/Frk Family

FRK is an NRTK Fyn-related kinase and is one of the members of the breast tumor kinase (BRK) family [166,167]. FRK is related to SFKs [166] and known as protein tyrosine kinase 5 (PTK5) that belongs to BRK family kinases (BFKs) [166]. FRK kinases include BRK and SRMS that contain C-terminal regulatory tyrosine and N-terminal myristylation sites [166,167,168]. In addition, the Frk is highly homologous to the Src family in terms of its sequence and structure [169]. Frk presents a highly divergent N-terminal sequence followed by an SH3 domain, an SH2 domain, and a tyrosine kinase domain [168,169]. Unlike the Src family, most of the Frk family members lack the N-myristoylation site [168,170,171], allowing them to be localized to the nucleus but not to the membrane [168].
The unique feature of FRK and IYKs is associated with the presence of a nuclear localization signal (NLS) within the SH2 domain [168,169,170,171,172]. This NLS is a bipartite motif that is organized into two clusters of basic amino acids that are separated by a spacer of nine amino acids [171,172]. The two NLS sequences of FRK are organized as follows: the first NLS sequence (KRLDEGGFFLTRRR) is followed by a spacer of nine amino acids extending from amino acid residue 168 to amino acid residue 181, followed by the second NLS sequence (KRLDEGGFFLTRRR) [172,173]. The two NLS sequences of IYK are organized as follows: the first NLS sequence (RRDEGGFFLTRRK) is followed by a spacer of nine amino acids extending from amino acid residue 175 to amino acid residue 188, followed by the second NLS sequence (RRDEGGFFLTRRK) [171,172,173].
SFKs are highly homologous in structure, consisting of four consecutive Src homology (SH4, HS3, SH2, and SH1) domains [168,174]. Of note, the SH4 domain serves as a membrane targeting region for myristylation and/or palmitoylation at the N-terminus of SFK members [175]. Thus, the myristylation and/or palmitoylation of SH4 is essential to facilitate the membrane localization of SKF members [176]. The structure of BRK/Frk family members is outlined in Figure 8.

2.8. Abl Family

The two Abl family members include Abl and Arg, which are widely expressed in all organs; while the highest expression level of these is noted in the thymus, spleen, and testes [177,178], the highest expression level of r expression Arg is observed in the brain [176,177]. The structure of Abl is like that of the Src family with an N-terminal region including one SH3, one SH2, and one tyrosine kinase domain [179,180]. The Abl family kinase lacks a C-terminal-negative regulatory site that is characteristic of Src family members [181,182] and instead possesses a C-terminal region that contains an F-actin-binding domain (FABD) [183,184,185,186]. The kinase domain is followed by a proline-rich DNA-binding domain [184,185,186]. The C-terminal region of Abl represents an actin-binding domain containing three nuclear localization signals (NLSs) and one nuclear export signal (NES) [187].
Arg shows high sequence similarity to Abl within the N-terminal region [188,189,190,191]. The sequence similarity of Arg within the C-terminal region is moderate compared to that of Abl [190,191]. However, the critical elements of Arg structure include the conserved SH3-binding sites and the actin-binding domain [192,193].
The mechanisms regulating the subcellular localization of Abl to the membrane are mediated by its own structural domains [194,195,196,197]. The common subcellular localization of Abl in different cell types localizes to the nucleus and cytoplasm, where they are mostly membranes and actin filaments [186,194,195,196,197]. The localization of Abl in hematopoietic cells and neurons is common in the cytoplasm, while the localization of Arg appears to only be cytoplasm [195,196,198]. The nuclear localization of Abl is determined by both NLS and NES sites [197], while its membrane localization is mediated via a myristoylation site-dependent mechanism [182,198], and the binding of Abl to the actin cytoskeleton is mediated via an actin-binding domain-dependent mechanism [182,199].
The activation of ABL family members (e.g., Abl1, ARG) is mediated by different mechanisms, including mutation, dimerization, and phosphorylation [200,201,202], and by the phosphorylation of the tyrosine 412 (Y412) and tyrosine 245 (Y245) residues within kinase and interlinked regions, respectively [203,204,205,206]. However, the constitutive activation of Abl by the disruption of its autoinhibition mediated by the translocation to a variety of many genes such as BCR, Tel, and ETV6 is the main cause of the development of different tumor types, including leukemia [207]. Although the role of both c-Abl and Arg in the development of a variety of human leukemias has been discussed, their oncogenic function in solid tumors has not been determined [208]. However, data from the immune histochemistry analysis of c-Abl and/or Arg show that both Abl and Arg are overexpressed in some solid tumors, including the brain, lung, ovarian, colon, and prostate cancer. In addition, c-Abl amplification was found in renal medulla carcinomas [209,210,211]. The structure and functional domains, as well as the 3D of ABL kinase protein, are outlined in Figure 9A,B.

2.9. Syk Kinase

The spleen tyrosine kinase (Syk) is localized mainly in different tissue types [15,212,213,214,215,216]. The members of Syk family kinases include Syk and Zap70, which are similar in terms of their structure, and they constitute two SH2 in tandem followed by a catalytic domain in the N-terminal half and catalytic kinase domain in the c-terminal region [217,218,219]. The members of the Syk family do not possess any fatty acid modification sites, and consequently, they are recognized as cytosolic proteins [220]. The localization of the members of the Syk family is dramatically affected by cell stimulation [221,222,223]. The SH2 tandem domains of the Syk family have been reported to exhibit extremely high affinity for doubly phosphorylated tyrosine-based immune receptor activation motifs (ITAMs) [217,223]. The aim of the phosphorylation of ITAMs by Src family PTK is to initiate signaling events for multiple receptors and subsequently generate an excellent binding platform for the Syk family [224,225]. Consequently, the cytosolic molecules of Zap and Syk can be translocated to the cytoplasmic tails of immune recognition receptors (MIRRs) in the proximity of their membrane [225,226]. The structural and functional domains of the Syk kinases are demonstrated (Figure 10).

3. Mechanisms of NRTK-Mediated Pathways in Normal and Cancer Cells

Unlike RTKs, which are transmembrane receptors, NRTKs are in the cytoplasm and play an important role in the regulation of various cellular functions through the modulation of intracellular signaling [11,12]. The NRTK family members are distributed among the different subcellular compartments [11,12] and are determined by either factor on the cell surface or by integrating heterologous protein–protein interactions in the cytoplasm [15,227].
As an intracellular cytoplasmic protein that transduces intracellular signaling, NRTKs have the potential to play an essential role in the modulation of signaling pathways such as the JAK pathway, which has the potential to regulate gene expression via the IL-6-mediated phosphorylation of membrane-bound tyrosine kinase (TK) and Janus kinase that is responsible for the activation of the signal transducer and activator of transcription (STAT) [15,228]. In addition to its role in the inhibition of cell growth and induction of apoptosis [13,18], Abl1 plays a central role in the regulation of tumor development and progression [229]. In human leukemias, the enhancement of the oncogenic activity of Abl is mediated by chromosomal translocation [t(9;22) (q34;q11)] to generate BCR-ABL1 fusion proteins, leading to the disruption of intramolecular interactions of inhibitory Abl1 [230]. The oncogenic activation of Abl1 in BCR-ABL1 chimeric proteins is mediated by an N-terminal coiled-coil (CC) oligomerization-dependent mechanism [226]. Also, the involvement of multiple signaling pathways such as RAS, NF-κB, PI3K-Akt, Jun, β-catenin, and STAT is considered [231].
The regulation of cellular functions in both normal and tumor cells by the members of NRTKs is common. For example, Fak has been reported to trigger the regulation of cell adhesion and proliferation [232,233], besides being one of the most important components of the signal transduction pathways of Fyn [234]. Also, the involvement of the Ack family kinases in the regulation of cell growth is associated with the induction of both JAK and Src-dependent mechanisms [235,236]. The Tec family of protein tyrosine kinases plays an important role in cellular signaling through an antigen-receptor-dependent mechanism such as the TCR, BCR, and FCε receptor [237,238]. Syk functions as a mediator for the immune response between cell receptors and intracellular signaling [218,220].
In contrast to the activation of RTKs that is determined by ligand-binding-dependent mechanisms, the activation of NRTKs is much more complex and is mediated by a heterologous protein–protein interaction, a mechanism that is essential to initiating signal transduction processes via a transphosphorylation-dependent mechanism [1,15,175].
The oncogenic activity of Abl1 in the context of the ABL1-BCR fusion protein is mediated via the BCR N-terminal coiled-coil (CC) oligomerization domain [208,239,240]. The activation of both c-Abl and Arg kinases in human primary melanomas is associated with melanoma cell invasion via a mechanism mediated by Stat3 [209,241]. Moreover, the co-activation of both c-Abl and Arg kinases in solid tumors was found to enhance platelet-derived growth factor (PDGF)-induced epithelial–mesenchymal transition and promote transforming growth factor-β (TGF-β)-induced tumor growth [15,211]. Also, the ability of Abl1 kinase to phosphorylate AKT and ERK and to promote cancer cell proliferation and survival has been reported [242].
The role of Abl1 in the development of HCC is mediated by the c-MYC/NOTCH1 axis [239]. Thus, the inhibition of Abl1 has been shown to be a promising therapeutic strategy to treat HCC in patients with overexpressed Abl1. Abl1 kinase inhibitors have been approved for their clinical relevance in many tumor types, including leukemia and HCC, in patients showing the overexpression and/or activation of Abl1 [243,244,245].
The JAK signal transducer and activator of the transcription (JAK-STAT) pathway are an evolutionarily conserved mechanism of transmembrane signal transduction that enables cells to communicate with their exterior environment [70,246]. The activation of the JAK-STAT pathway is regulated by various factors, including cytokines, interferons, and growth factors [70,247]. Consequently, the JAK-STAT pathway has the potential to drive physiological and pathological processes, leading to the regulation of cell proliferation and metabolism in addition to inducing immune responses and enhancing inflammation and malignancy [70,247]. Thus, the dysregulation of JAK-STAT signaling in combination with related genetic mutations is strongly associated with immune activation and tumor progression [248,249].
Fes kinase is characterized by its oncogenic properties that have been confirmed both in cell lines and in animal models [250,251]. FES kinase has a limited expression pattern in both endothelial and epithelial cells, particularly those of neuronal origin, and in cells with hematopoietic origin, such as myeloid cells [15,250,251]. The therapeutic impact of Fes in tumor therapy relies on evidence suggesting the role Fes kinase in the regulation of tumor microenvironment-dependent mechanisms and its functional role as a tumor suppressor protein in melanoma cell lines [15,252]. Moreover, targeting Fes-related kinase (Fer) was found to have an inhibitory effect on melanoma growth and metastasis [253,254]. Beyond its impact as a therapeutic target in melanoma, the overexpression of Fer is correlated with cancer cell survival in non-small-cell lung cancer (NSCLC) as well as in lung adenocarcinoma cell invasion and tumor metastasis, besides being a prognostic marker for poor prognosis [255,256]. The loss of Fer has been shown to impair metabolic plasticity both in lung and breast carcinoma cells [257]. Also, the role of Fer kinase in the regulation of the mechanisms of HCC metastatic dissemination has been reported [258].
The involvement of integrins in the regulation of many fundamental cellular processes, including cell adhesion and migration, has been widely reported [248,249]. As transmembrane adhesion receptors, integrins are localized at cell–matrix contact sites, where they connect components of the extracellular matrix (ECM) and interact with several structural and signaling molecules, such as talin, paxillin, vinculin, ά-actin, FAK, and Src [248,249]. Thus, as integrin mediators, both FAK and Src kinases have the potential to have cell adhesion-dependent responses by direct and indirect interaction with integrins [250,251]. The direct interaction of ß3 integrins via their cytoplasmic tail with the SH3 domain of Src kinase results in the activation of Src that in turn triggers the activation of the guanine nucleotide exchange factor (GEF) and T cell lymphoma invasion and metastasis 1 (Tiam1) [250,251,252]. These factors together have been reported to be essential for the activation of Rac that can then induce the stimulation of actin-driven positive activity at the site of integrin binding, finally leading to cell adhesion or migration [250,251,252].
Also, the interaction of ß3 integrin with the SH3 domain has been shown to trigger the stimulation of STAT3 and FAK signaling, leading to tumor growth [253,254]. The phosphorylation of ß1 integrin is required to inhibit Rho-mediated cytoskeletal contractility in addition to mediating the regulation phenotype transformation [253,254]. Aside from its contribution in the modulation of several cellular functions, Src can also weaken the connections between ECM, integrins, and cytoskeleton, leading to cell adhesion turnover and remodeling of the actin cytoskeleton, suggesting an important role for FAK in the modulation of integrin signaling [253,254,255,256]. For example, FAK has been reported to act as a signaling molecule and scaffold and recruit Src and Src substrates to integrin sites [255,256,257,258]. Consequently, the cell–matrix contact-dependent clustering of integrins is expected to be the main mechanism leading to the activation of FAK, which has been reported to result from the cell–matrix contact-dependent clustering of integrins [259,260].

4. NRTKs as Therapeutic Target in Tumor Treatment

NRTKs are key players in intracellular signaling pathways, and their dysregulation in the form of constitutive activation is expected to be associated with tumor initiation, progression, and drug resistance. Thus, targeting aberrantly activated NRTKs may prevent tumorigenesis in addition to having an impact as a therapeutic target for anticancer agents.
The analysis of BRAF mutant melanomas revealed a contributing role for Abl kinase activity in the enhancement of acquired resistance to BRAF and MEK inhibitors in therapy-resistant cells [261]. Abl kinases are expected to play an important role in the development of drug resistance in cancer since the suppression of Abl1/Abl2 activity by selective kinase inhibitors Nilotinib or genetic inhibitors was found to sensitize cancer cells to BRAF and MEK inhibitors [257]. Some clinically available tyrosine kinase inhibitors, particularly those targeting Abl kinases, have been approved in BCR-Abl-driven chronic myelogenous leukemia (CML) [262], as well as in many solid tumors, including lung cancer [263,264].
As such, Fes has been suggested to exert both tumor-stimulatory and tumor-suppressive effects on cancer cells based on the type of malignancy [265,266,267,268].
Although Fes represents a potential therapeutic target in different tumor types, its inhibition in macrophage and vascular endothelial cells attenuates their tumor-promoting roles [269,270,271,272].
JAK/STAT is one of the versatile signaling pathways whose crucial role in tumor progression has been extensively studied. The potential interaction of JAK/STAT with many signaling pathways suggested it as a potential therapeutic target for the development of clinically relevant small molecular inhibitors. The abnormal activation of JAK/STAT signaling is frequently common in a variety of different tumor types.
Activation of the JAK-STAT pathway is mediated by the ligation of growth factor receptors with the corresponding ligands, leading to receptor dimerization and the recruitment of related JAKs. Activated JAK then becomes able to phosphorylate tyrosine residues of the receptors to form docking sites for the phosphorylation of STATs and other substrates to trigger abnormal cell proliferation tumor initiation, progression, and drug resistance [273,274,275,276]. In skin cancer, particularly melanoma, the activation of the JAK/Stat pathway is mostly associated with tumor progression and treatment resistance [277,278]. Melanoma has been reported to display distinct gene expression profiles that depend on the activation levels of individual STAT proteins [279]. Other studies indicate that the phosphorylation of STATs may function as a predictive marker for melanoma progression [278,280]. The alteration of STAT proteins both in melanoma cells and immune cells of melanoma patients has been reported [281,282]. Thus, targeting the JAK/STAT pathway is expected to have a therapeutic impact on melanoma treatment [283,284].
In addition to their involvement in various signaling networks, the JAK/STAT pathway is continuously in crosstalk with different pathways, particularly those associated with hematological malignancies and melanoma [285,286].
The evaluation of the JAK/STAT pathway as a marker for esophageal squamous cell carcinoma (ESCC) progression has been confirmed in tumors and adjacent normal esophageal epithelia by immunohistochemistry [286]. Also, the increased activation of p-JAK1 and p-STAT3 in primary tumors revealed their reliability as a predictive marker for poor prognosis in ESCC patients [287]. Thus, beside serving as a predictive marker with a poor prognosis and playing a crucial role in the regulation of ESCC progression, the JAK/STAT pathway is a promising therapeutic target for tumor treatment [288,289]. Targeting JAK2 by its specific inhibitor, AG490, was found to decrease the inflammation and development of ESCC via the mechanism mediated by the suppression of STAT3 [290,291].
In HCC, the phosphorylation of STAT3 has been reported to target growth factors and cytokines like TNF, IL-6, hepatocytes growth factor (HGF), and epidermal growth factor (EGF) family [292,293]. However, the aberrant activation of SAT3 in response to the mentioned stimulators has been shown to be a JAK1-depnednt mechanism and is associated with HCC progression and metastasis [294,295].
The tumorigenic role of JAK/STAT signaling has been reported in different tumors including head and neck squamous cell carcinoma (HNSCC) [296,297] and in breast cancer [298]. In addition, JAKs and STATs were found to contribute to the development of breast cancer by acting as oncogenes [299,300]. Analysis of breast cancer specimens revealed that JAK1, JAK2, and JAK3 are frequently targeted for somatic mutations, which may be the main cause of the development of breast cancer [301,302,303].
Ack1 has been shown to have a therapeutic impact on various tumor types, including breast cancer [304], HCC [305,306,307], and colorectal cancer [308,309,310]. The inhibition of Ack1 alone was found to inhibit proliferation and migration in many tumor types [305,306,307] and to enhance the apoptosis of lung cancer cells [311].
Syk is a modulator of tumorigenesis that can drive opposite functions. In addition to having tumorigenic activity in some cells, Syk can also act as a tumor suppressor in other tumor cells [312]. In hematopoietic malignancies, Syk is mostly associated with tumor progression and survival, while its inhibition or silencing is frequently associated with the induction of apoptosis [313,314,315,316,317]. In cancers of non-immune cells, in addition to initiating a pro-survival signal, Syk suppresses tumorigenesis via mechanisms mediated by the restriction of epithelial–mesenchymal transition, improvement in cell–cell interactions, and inhibition of migration [318,319].
The activation of Syk is mediated by the binding of its SH2 domains to the immunoreceptor tyrosine-based activation motif (ITAM) that is known as a common sequence of amino acids located on most SYK-coupled receptors [320,321].
Thus, targeting the Syk signaling pathway is a promising therapeutic target for diseases such as cancer. Entospletinib, for example, is one Syk inhibitor that has shown promising therapeutic effects against B cell malignancies in clinical trials [322]. In addition, some combinational regimens based on Syk inhibitors have been evaluated in several malignancies [323,324,325].
IL-2-inducible kinase (ITK) is a member of the Tec kinase family that is essential in driving T cell receptor (TCR) signaling [326]. ITK kinase has been reported to play an essential role in the development of acute lymphoblastic T cell leukemia and cutaneous T cell lymphoma [327]. Thus, in addition to being a promising therapeutic target for other human diseases, ITK kinase has been approved as an effective therapeutic target for T cell malignant lymphoma [327].
Targeting Fak was found to enhance the therapeutic efficacy of both radio- and chemotherapy-based treatments [324,328]. Most chemotherapy drugs, particular platinum-based agents and fluoropyrimidine chemotherapy as well as radiation, mediate their anticancer effects in cancer cells via DNA damage-dependent mechanisms [329,330].
In addition to its crucial role in the regulation of DNA damage and repair mechanisms, Fak can also coordinate the nuclear translocation of β-catenin to enhance the transcription of genes involved in the regulation of DNA repair mechanisms to promote cell survival and drug resistance [331,332]. Therefore, targeting FAK is a promising strategy to restore sensitivity to DNA damage therapy [324,333].
The role of Src kinase in the regulation of tumor growth and metastasis has been reported [334,335]. The involvement of Src in the regulation of tumor cell-induced bone destruction has been shown to be an osteoclasts-dependent mechanism since Src has been reported to enhance osteoclast activation and bone resorption, mechanisms that are thought to be associated with the progression of bone metastases [336].
Src inhibitors have been approved for their ant tumor activities in different tumor types, including osteoporosis and osteolytic bone disease [337]. The most common Src inhibitors include NVP-AAK980, an ATP target compound [236], AZD0530 (Saracatinib) [338], AP23451 [339], and the multi-targeted kinase inhibitor Dasatinib (BMS-354825). These can all effectively inhibit ABL-BCR, Src family, and other kinases via a mechanism mediated through the destruction of the ATP-binding site of the enzymes [340]. However, the most used clinical inhibitors are Saracatinib and Dasatinib inhibitors [341].
Csk is the member of the SFK family and has the potential to mediate its inhibitory effects on SFK activity by the phosphorylation of its C-terminal Tyr527 residue of the negative regulatory loop [342]. Csk has the potential to trigger tumor suppression through the inhibition of the oncogenic activity of SFKs [342]. The overexpression of Csk has been shown to inhibit the growth of human colon cancer cells [343]. In addition to its inhibitory effect on tumor growth, the loss of Csk was found to be associated with the development of HCC in a murine model [344].
Several kinase inhibitors targeting family members have been further developed and approved for their clinical application. The most common clinically applied inhibitors include imatinib/(STI-571), Dasatinib/(BMS-354825, Nilotinib/(AMN107), Bosutinib (SKI-606), Radotinib6/(IY-5511), Ponatinib/(AP24534), Asciminib/(ABL001), Ibrutinib/(PCI-32765), Acalabrutinib/(ACP-196), Zanubrutinib/(BGB-3111), Ruxolitinib/(INC424), and Fedratinib/(SAR302503, TG101348).
Imatinib is an Abl tyrosine kinase inhibitor (TKI) with high affinity to bind to the ATP-binding pocket of the Abl kinase domain and has the potential to BCR-ABL kinase [341]. Therefore, a mutation in the threonine 315 residue (T315) of the ATP-binding pocket is critical for the accessibility of the ATP-binding pocket and can confer resistance to imatinib and the related TKIs [341]. Second-generation BCR-ABL inhibitors such as Nilotinib/(AMN107) have a clinical utilization which has confirmed the therapeutic potential of second-generation TKIs to overcome imatinib resistance caused by ABL kinase point mutations [341]. Of note, Nilotinib is an ATP-competitive type II kinase inhibitor with a greatly improved potency compared to imatinib [341,342].
Bosutinib belongs to the second-generation of TKIs besides being characterized with its dual inhibitory activity against SRC and ABL kinases [341,342,343,344]. Bosutinib inhibitor has minimal inhibitory effects against other tyrosine kinases such as c-KIT and PDGFR-β kinases, but it is more effective against imatinib-resistant leukemias [341,342,343,344]. The inhibitory effect of Bosutinib was more pronounced in patients with Philadelphia chromosome-positive chronic myeloid leukemia (Ph+ CML), particularly in patients who are resistant to treatment with imatinib [345]. Other agents such as Radotinib, that is clinically approved as a second-generation BCR-ABL inhibitor, can exhibit inhibitory effects on some imatinib-resistant tumors [346]. Also, the allosteric inhibitor Asciminib has been reported for its affinity to bind to the myristate pocket of BCRABL and is thereby effective against T315I-mutant BCR-ABL [347]. The third-generation inhibitor, Ponatinib, is clinically approved for the treatment of both wild-type and T315I-mutant BCR-ABL [341,348].
Ponatinib is a third generation of TKI that is characterized by its inhibitory effects on multiple kinases besides being an effective BCR-ABL1 inhibitor with higher activity against the ABL kinase domain with mutations such as tyrosine315 isoleucine mutation (T315I) [348,349]. The antitumor activity of Ponatinib as TKI relies on its ability to inhibit the catalytic activity of native and mutant ABL [350,351]. In addition to the unique, multi-targeted properties of Ponatinib, a number of studies have demonstrated its ability to target other important tyrosine kinases (FGFR, PDGFR, SRC, RET, KIT, and FLT1) in different human malignancies [350,351].
Thus, based on its functional role in tumor progression SFK is a potential therapeutic target [351]. The potential of Dasatinib to target multiple kinases, including Src, Fgr, Fyn, Hck, Lck, Lyn, and Ye, has been widely documented [341,350,351].

5. Non-Receptor Tyrosine Kinases and Protein Tyrosine Phosphatases

The regulation of the activity of NRTKs is mediated by protein tyrosine phosphatases (PTPs), leading to the dephosphorylation of protein targets and adopting the activity of protein kinases in response to physiological condition [11,352]. PTPs play an important role in the regulation of tyrosine phosphorylation and signal transduction [353,354].
These PTPs consist of functionally conserved PTPase domains with highly conserved signature motifs (His/Val) Cys(X)5Arg (Ser/Thr), which are characterized by their catalytic function, leading to hydrolyzing phosphate from the tyrosine residues of their substrates [5]. PTPs are tyrosine-specific phosphatases or dual-specific phosphatases [355,356]. The main function of specific PTPs is to dephosphorylate only tyrosine residues [355,356], while the function of dual-specific PTPs is to dephosphorylate both tyrosine and serine/threonine residues [355,356]. However, cellular localization PTPs can also be classified into receptor or non-receptor PTPs receptor types referred to PTPs that span the cell membrane, while non-receptor PTPs are referred to PTPs that are localized mainly in the cytoplasm [357,358]. In addition to their abovementioned classifications, PTPs can be classified into tumor-suppressive or oncogenic PTPs based on the type of dephosphorylated protein kinase of interest [359,360]. Src homology-2-containing protein tyrosine phosphatase 2 (SHP-2) is a non-receptor, tyrosine-specific PTP that binds to the intracellular domain of RTKs via phosphotyrosine residues [361,362,363]. SHP-2 can dephosphorylate proteins or residues that are able to suppress the function of Ras and SFK inhibitors to enhance the activity of tumorigenic pathways such as PI3K and MAPK/ERK signaling [363,364].

6. Conclusions

Based on their profound contribution to the regulation of numerous vital cellular mechanisms, non-receptor tyrosine kinases (NRTKs) play an essential role in tumor development, progression, and treatment resistance. Thus, targeting NRTKs is a relevant strategy that can prevent tumorigenesis and its consequences. Also, the clinical use of tyrosine kinase inhibitors, in combination with conventional treatments, allows for targeted-based cancer therapies using small specific inhibitors, which are less toxic than traditional cytotoxic chemotherapy. Therefore, the establishment of effective therapeutic strategies based on the dysregulation of NRTK molecular functioning is essential for patient care.

Author Contributions

Conceptualization, M.H., A.M.E., T.W.F., S.S., S.-Y.H., S.-L.H. and Y.H.; methodology, M.H., A.M.E., T.W.F., S.A.A.A., W.A.A.-A., S.S., S.-Y.H., S.-L.H., M.L.B. and M.M.; software, M.H., A.M.E., T.W.F., S.S., S.A.A.A., W.A.A.-A., M.O.A. (Maryam O. Alamodi), M.O.A. (Mohammed O. Alamodi), S.-Y.H., S.-L.H., M.L.B. and M.M.; validation, M.H., A.M.E., T.W.F., S.S., S.-Y.H., S.-L.H., M.L.B. and M.M.; formal analysis, M.H., A.M.E., T.W.F., S.S., S.-Y.H., S.A.A.A., W.A.A.-A., S.-L.H., M.L.B. and M.M.; investigation, M.H., A.M.E., T.W.F., S.S., S.-Y.H., S.-L.H., M.L.B. and M.M.; resources, M.H., A.M.E., T.W.F., S.S. and S.-Y.H.; data curation, M.H., A.M.E. and T.W.F.; writing—original draft preparation, M.H., A.M.E., T.W.F., S.S., S.-Y.H., M.L.B. and M.M.; writing—review and editing, M.H., A.M.E., T.W.F., S.S., S.-Y.H., M.L.B. and M.M.; visualization, M.H., A.M.E., T.W.F., S.S., S.-Y.H., S.-L.H., M.L.B. and M.M.; supervision, M.H.; project administration, M.H.; funding acquisition, M.H. and A.M.E. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors have no conflicts of interest to disclose.

References

  1. Bhanumathy, K.; Balagopal, A.; Vizeacoumar, F.S.; Vizeacoumar, F.J.; Freywald, A.; Giambra, V. Protein Tyrosine Kinases: Their Roles and Their Targeting in Leukemia. Cancers 2021, 13, 184. [Google Scholar] [CrossRef] [PubMed]
  2. Díaz Galicia, M.E.; Aldehaiman, A.; Hong, S.; Arold, S.T.; Grünberg, R. Methods for the recombinant expression of active tyrosine kinase domains: Guidelines and pitfalls. Methods Enzym. 2019, 621, 131–152. [Google Scholar] [CrossRef]
  3. Kim, M.; Baek, M.; Kim, D.J. Protein Tyrosine Signaling and its Potential Therapeutic Implications in Carcinogenesis. Curr. Pharm. Des. 2017, 23, 4226–4246. [Google Scholar] [CrossRef] [PubMed]
  4. Aschner, Y.; Downey, G.P. The Importance of Tyrosine Phosphorylation Control of Cellular Signaling Pathways in Respiratory Disease: pY and pY Not. Am. J. Respir. Cell Mol. Biol. 2018, 59, 535–547. [Google Scholar] [CrossRef] [PubMed]
  5. Tautz, L.; Critton, D.A.; Grotegut, S. Protein tyrosine phosphatases: Structure, function, and implication in human disease. Methods Mol. Biol. 2013, 1053, 179–221. [Google Scholar] [CrossRef] [PubMed]
  6. Ghoreschi, K.; Laurence, A.; O’Shea, J.J. Janus kinases in immune cell signaling. Immunol. Rev. 2009, 228, 273–287. [Google Scholar] [CrossRef] [PubMed]
  7. Navarro, M.N.; Cantrell, D.A. Serine-threonine kinases in TCR signaling. Nat. Immunol. 2014, 15, 808–814. [Google Scholar] [CrossRef]
  8. Lease, K.A.; Lau, N.Y.; Schuster, R.A.; Torii, K.U.; Walker, J.C. Receptor serine/threonine protein kinases in signalling: Analysis of the erecta receptor-like kinase of Arabidopsis thaliana. New Phytol. 2001, 151, 133–143. [Google Scholar] [CrossRef] [PubMed]
  9. Paul, M.K.; Mukhopadhyay, A.K. Tyrosine kinase-Role and significance in Cancer. Int. J. Med. Sci. 2004, 1, 101–115. [Google Scholar] [CrossRef]
  10. Anreddy, N.; Gupta, P.; Kathawala, R.J.; Patel, A.; Wurpel, J.N.; Chen, Z.S. Tyrosine kinase inhibitors as reversal agents for ABC transporter mediated drug resistance. Molecules 2014, 19, 13848–13877. [Google Scholar] [CrossRef]
  11. Solouki, S.; August, A.; Huang, W. Non-receptor tyrosine kinase signaling in autoimmunity and therapeutic implications. Pharmacol. Ther. 2019, 201, 39–50. [Google Scholar] [CrossRef]
  12. Gocek, E.; Moulas, A.N.; Studzinski, G.P. Non-receptor protein tyrosine kinases signaling pathways in normal and cancer cells. Crit. Rev. Clin. Lab. Sci. 2014, 51, 125–137. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, J.Q.; Derges, J.D.; Bodepudi, A.; Pokala, N.; Mao, L.M. Roles of non-receptor tyrosine kinases in pathogenesis and treatment of depression. J. Integr. Neurosci. 2022, 21, 25. [Google Scholar] [CrossRef] [PubMed]
  14. Schlessinger, J. Receptor tyrosine kinases: Legacy of the first two decades. Cold Spring Harb. Perspect. Biol. 2014, 6, a008912. [Google Scholar] [CrossRef] [PubMed]
  15. Siveen, K.S.; Prabhu, K.S.; Achkar, I.W.; Kuttikrishnan, S.; Shyam, S.; Khan, A.Q.; Merhi, M.; Dermime, S.; Uddin, S. Role of Non Receptor Tyrosine Kinases in Hematological Malignances and its Targeting by Natural Products. Mol. Cancer 2018, 17, 31. [Google Scholar] [CrossRef] [PubMed]
  16. Li, E.; Hristova, K. Role of receptor tyrosine kinase transmembrane domains in cell signaling and human pathologies. Biochemistry 2006, 45, 6241–6251. [Google Scholar] [CrossRef] [PubMed]
  17. Benati, D.; Baldari, C.T. SRC family kinases as potential therapeutic targets for malignancies and immunological disorders. Curr. Med. Chem. 2008, 15, 1154–1165. [Google Scholar] [CrossRef] [PubMed]
  18. Hubbard, S.R.; Miller, W.T. Receptor tyrosine kinases: Mechanisms of activation and signaling. Curr. Opin. Cell Biol. 2007, 19, 117–123. [Google Scholar] [CrossRef] [PubMed]
  19. Kan, Y.; Paung, Y.; Seeliger, M.A.; Miller, W.T. Domain Architecture of the Nonreceptor Tyrosine Kinase Ack1. Cells 2023, 12, 900. [Google Scholar] [CrossRef]
  20. Seok, S.H. Structural Insights into Protein Regulation by Phosphorylation and Substrate Recognition of Protein Kinases/Phosphatases. Life 2021, 11, 957. [Google Scholar] [CrossRef]
  21. Krupa, A.; Srinivasan, N. Diversity in domain architectures of Ser/Thr kinases and their homologues in prokaryotes. BMC Genom. 2005, 6, 129. [Google Scholar] [CrossRef] [PubMed]
  22. Prieto-Echagüe, V.; Gucwa, A.; Craddock, B.P.; Brown, D.A.; Miller, W.T. Cancer-associated mutations activate the nonreceptor tyrosine kinase Ack1. J. Biol. Chem. 2010, 285, 10605–10615. [Google Scholar] [CrossRef] [PubMed]
  23. Hubbard, S.R.; Till, J.H. Protein tyrosine kinase structure and function. Annu. Rev. Biochem. 2000, 69, 373–398. [Google Scholar] [CrossRef] [PubMed]
  24. Sun, G.; Ayrapetov, M.K. Dissection of the catalytic and regulatory structure-function relationships of Csk protein tyrosine kinase. Front. Cell Dev. Biol. 2023, 11, 1148352. [Google Scholar] [CrossRef] [PubMed]
  25. Gan, W.; Roux, B. Binding specificity of SH2 domains: Insight from free energy simulations. Proteins 2009, 74, 996–1007. [Google Scholar] [CrossRef] [PubMed]
  26. Pawson, T.; Gish, G.D.; Nash, P. SH2 domains, interaction modules and cellular wiring. Trends Cell Biol. 2001, 11, 504–511. [Google Scholar] [CrossRef] [PubMed]
  27. Mahajan, K.; Mahajan, N.P. ACK1/TNK2 tyrosine kinase: Molecular signaling and evolving role in cancers. Oncogene 2015, 34, 4162–4167. [Google Scholar] [CrossRef] [PubMed]
  28. Yokoyama, N.; Miller, W.T. Biochemical properties of the Cdc42-associated tyrosine kinase ACK1. Substrate specificity, authphosphorylation, and interaction with Hck. J. Biol. Chem. 2003, 278, 47713–47723. [Google Scholar] [CrossRef] [PubMed]
  29. Yang, W.; Cerione, R.A. Cloning and characterization of a novel Cdc42-associated tyrosine kinase, ACK-2, from bovine brain. J. Biol. Chem. 1997, 272, 24819–24824. [Google Scholar] [CrossRef]
  30. Prieto-Echagüe, V.; Miller, W.T. Regulation of ack-family nonreceptor tyrosine kinases. J. Signal Transduct. 2011, 2011, 742372. [Google Scholar] [CrossRef]
  31. Gajiwala, K.S.; Maegley, K.; Ferre, R.; He, Y.A.; Yu, X. Ack1: Activation and regulation by allostery. PLoS ONE 2013, 8, e53994. [Google Scholar] [CrossRef]
  32. Ahmed, S.; Miller, W.T. The noncatalytic regions of the tyrosine kinase Tnk1 are important for activity and substrate specificity. J. Biol. Chem. 2022, 298, 102664. [Google Scholar] [CrossRef]
  33. Umarao, P.; Rath, P.P.; Gourinath, S. Cdc42/Rac Interactive Binding Containing Effector Proteins in Unicellular Protozoans with Reference to Human Host: Locks of the Rho Signaling. Front. Genet. 2022, 13, 781885. [Google Scholar] [CrossRef]
  34. Pao-Chun, L.; Chan, P.M.; Chan, W.; Manser, E. Cytoplasmic ACK1 interaction with multiple receptor tyrosine kinases is mediated by Grb2: An analysis of ACK1 effects on Axl signaling. J. Biol. Chem. 2009, 284, 34954–34963. [Google Scholar] [CrossRef] [PubMed]
  35. Hayashi, S.Y.; Craddock, B.P.; Miller, W.T. Phosphorylation of Ack1 by the Receptor Tyrosine Kinase Mer. Kinases Phosphatases 2023, 1, 167–180. [Google Scholar] [CrossRef] [PubMed]
  36. Chan, W.; Tian, R.; Lee, Y.F.; Sit, S.T.; Lim, L.; Manser, E. Down-regulation of active ACK1 is mediated by association with the E3 ubiquitin ligase Nedd4-2. J. Biol. Chem. 2009, 284, 8185–8194. [Google Scholar] [CrossRef] [PubMed]
  37. Lougheed, J.C.; Chen, R.H.; Mak, P.; Stout, T.J. Crystal structures of the phosphorylated and unphosphorylated kinase domains of the Cdc42-associated tyrosine kinase ACK1. J. Biol. Chem. 2004, 279, 44039–44045. [Google Scholar] [CrossRef]
  38. Yang, W.; Lo, C.G.; Dispenza, T.; Cerione, R.A. The Cdc42 target ACK2 directly interacts with clathrin and influences clathrin assembly. J. Biol. Chem. 2001, 276, 17468–17473. [Google Scholar] [CrossRef]
  39. Hodder, S.; Fox, M.; Binti Ahmad Mokhtar, A.M.; Mott, H.R.; Owen, D. ACKnowledging the role of the Activated-Cdc42 associated kinase (ACK) in regulating protein stability in cancer. Small GTPases 2023, 14, 14–25. [Google Scholar] [CrossRef]
  40. Galisteo, M.L.; Yang, Y.; Ureña, J.; Schlessinger, J. Activation of the nonreceptor protein tyrosine kinase Ack by multiple extracellular stimuli. Proc. Natl. Acad. Sci. USA 2006, 103, 9796–9801. [Google Scholar] [CrossRef]
  41. Linseman, D.A.; Heidenreich, K.A.; Fisher, S.K. Stimulation of M3 muscarinic receptors induces phosphorylation of the Cdc42 effector activated Cdc42Hs-associated kinase-1 via a Fyn tyrosine kinase signaling pathway. J. Biol. Chem. 2001, 276, 5622–5628. [Google Scholar] [CrossRef] [PubMed]
  42. La Torre, A.; del Mar Masdeu, M.; Cotrufo, T.; Moubarak, R.S.; del Río, J.A.; Comella, J.X.; Soriano, E.; Ureña, J.M. A role for the tyrosine kinase ACK1 in neurotrophin signaling and neuronal extension and branching. Cell Death Dis. 2013, 4, e602. [Google Scholar] [CrossRef] [PubMed]
  43. Mahajan, N.P.; Whang, Y.E.; Mohler, J.L.; Earp, H.S. Activated tyrosine kinase Ack1 promotes prostate tumorigenesis: Role of Ack1 in polyubiquitination of tumor suppressor Wwox. Cancer Res. 2005, 65, 10514–10523. [Google Scholar] [CrossRef] [PubMed]
  44. Van der Horst, E.H.; Degenhardt, Y.Y.; Strelow, A.; Slavin, A.; Chinn, L.; Orf, J.; Rong, M.; Li, S.; See, L.H.; Nguyen, K.Q.; et al. Metastatic properties and genomic amplification of the tyrosine kinase gene ACK1. Proc. Natl. Acad. Sci. USA 2005, 102, 15901–15906. [Google Scholar] [CrossRef] [PubMed]
  45. Kim, E.H.; Cao, D.; Mahajan, N.P.; Andriole, G.L.; Mahajan, K. ACK1-AR and AR-HOXB13 signaling axes: Epigenetic regulation of lethal prostate cancers. NAR Cancer 2020, 2, zcaa018. [Google Scholar] [CrossRef] [PubMed]
  46. Mahajan, K.; Coppola, D.; Challa, S.; Fang, B.; Chen, Y.A.; Zhu, W.; Lopez, A.S.; Koomen, J.; Engelman, R.W.; Rivera, C.; et al. Ack1 mediated AKT/PKB tyrosine 176 phosphorylation regulates its activation. PLoS ONE 2010, 5, e9646. [Google Scholar] [CrossRef] [PubMed]
  47. Shen, F.; Lin, Q.; Gu, Y.; Childress, C.; Yang, W. Activated Cdc42-associated kinase 1 is a component of EGF receptor signaling complex and regulates EGF receptor degradation. Mol. Biol. Cell 2007, 18, 732–742. [Google Scholar] [CrossRef] [PubMed]
  48. Howlin, J.; Rosenkvist, J.; Andersson, T. TNK2 preserves epidermal growth factor receptor expression on the cell surface and enhances migration and invasion of human breast cancer cells. Breast Cancer Res. 2008, 10, R36. [Google Scholar] [CrossRef]
  49. Xie, B.; Zen, Q.; Wang, X.; He, X.; Xie, Y.; Zhang, Z.; Li, H. ACK1 promotes hepatocellular carcinoma progression via downregulating WWOX and activating AKT signaling. Int. J. Oncol. 2015, 46, 2057–2066. [Google Scholar] [CrossRef]
  50. Angappulige, D.H.; Mahajan, N.P.; Mahajan, K. Epigenetic underpinnings of tumor-immune dynamics in prostate cancer immune suppression. Trends Cancer 2024, 10, 369–381. [Google Scholar] [CrossRef]
  51. Kan, Y.; Miller, W.T. Activity of the nonreceptor tyrosine kinase Ack1 is regulated by tyrosine phosphorylation of its Mig6 homology region. FEBS Lett. 2022, 596, 2808–2820. [Google Scholar] [CrossRef] [PubMed]
  52. Shinmura, K.; Kiyose, S.; Nagura, K.; Igarashi, H.; Inoue, Y.; Nakamura, S.; Maeda, M.; Baba, M.; Konno, H.; Sugimura, H. TNK2 gene amplification is a novel predictor of a poor prognosis in patients with gastric cancer. J. Surg. Oncol. 2014, 109, 189–197. [Google Scholar] [CrossRef] [PubMed]
  53. Fox, M.; Crafter, C.; Owen, D. The non-receptor tyrosine kinase ACK: Regulatory mechanisms, signalling pathways and opportunities for attACKing cancer. Biochem. Soc. Trans. 2019, 47, 1715–1731. [Google Scholar] [CrossRef]
  54. Chua, B.T.; Lim, S.J.; Tham, S.C.; Poh, W.J.; Ullrich, A. Somatic mutation in the ACK1 ubiquitin association domain enhances oncogenic signaling through EGFR regulation in renal cancer derived cells. Mol. Oncol. 2010, 4, 323–334. [Google Scholar] [CrossRef]
  55. Gu, J.; Qian, L.; Zhang, G.; Mahajan, N.P.; Owonikoko, T.K.; Ramalingam, S.S.; Sun, S.Y. Inhibition of ACK1 delays and overcomes acquired resistance of EGFR mutant NSCLC cells to the third generation EGFR inhibitor, osimertinib. Lung Cancer 2020, 150, 26–35. [Google Scholar] [CrossRef]
  56. Jones, S.; Cunningham, D.L.; Rappoport, J.Z.; Heath, J.K. The non-receptor tyrosine kinase Ack1 regulates the fate of activated EGFR by inducing trafficking to the p62/NBR1 pre-autophagosome. J. Cell Sci. 2014, 127, 994–1006. [Google Scholar] [CrossRef] [PubMed]
  57. Xu, S.H.; Huang, J.Z.; Chen, M.; Zeng, M.; Zou, F.Y.; Chen, D.; Yan, G.R. Amplification of ACK1 promotes gastric tumorigenesis via ECD-dependent p53 ubiquitination degradation. Oncotarget 2017, 8, 12705–12716. [Google Scholar] [CrossRef]
  58. Mahajan, K.; Lawrence, H.R.; Lawrence, N.J.; Mahajan, N.P. ACK1 tyrosine kinase interacts with histone demethylase KDM3A to regulate the mammary tumor oncogene HOXA1. J. Biol. Chem. 2014, 289, 28179–28191. [Google Scholar] [CrossRef]
  59. Sawant, M.; Wilson, A.; Sridaran, D.; Mahajan, K.; O’Conor, C.J.; Hagemann, I.S.; Luo, J.; Weimholt, C.; Li, T.; Roa, J.C.; et al. Epigenetic reprogramming of cell cycle genes by ACK1 promotes breast cancer resistance to CDK4/6 inhibitor. Oncogene 2023, 42, 2263–2277. [Google Scholar] [CrossRef]
  60. Sridaran, D.; Chouhan, S.; Mahajan, K.; Renganathan, A.; Weimholt, C.; Bhagwat, S.; Reimers, M.; Kim, E.H.; Thakur, M.K.; Saeed, M.A.; et al. Inhibiting ACK1-mediated phosphorylation of C-terminal Src kinase counteracts prostate cancer immune checkpoint blockade resistance. Nat. Commun. 2022, 13, 6929. [Google Scholar] [CrossRef]
  61. He, W.; Xu, L.; Ding, J.; Song, L.; Yang, W.; Klooster, I.; Pilco-Janeta, D.F.; Serrano, C.; Fang, H.; Jiang, G.; et al. Co-targeting of ACK1 and KIT triggers additive anti-proliferative and -migration effects in imatinib-resistant gastrointestinal stromal tumors. Biochim. Biophys. Acta Mol. Basis Dis. 2023, 1869, 166690. [Google Scholar] [CrossRef] [PubMed]
  62. Raheem, F.; Karikalan, S.A.; Batalini, F.; El Masry, A.; Mina, L. Metastatic ER+ Breast Cancer: Mechanisms of Resistance and Future Therapeutic Approaches. Int. J. Mol. Sci. 2023, 24, 16198. [Google Scholar] [CrossRef] [PubMed]
  63. Kawiak, A.; Kostecka, A. Regulation of Bcl-2 Family Proteins in Estrogen Receptor-Positive Breast Cancer and Their Implications in Endocrine Therapy. Cancers 2022, 14, 279. [Google Scholar] [CrossRef] [PubMed]
  64. Ji, Z.; Njauw, C.N.; Guhan, S.; Kumar, R.; Reddy, B.; Rajadurai, A.; Flaherty, K.; Tsao, H. Loss of ACK1 Upregulates EGFR and Mediates Resistance to BRAF Inhibition. J. Investig. Dermatol. 2021, 141, 1317–1324.e1311. [Google Scholar] [CrossRef] [PubMed]
  65. Zhu, L.; Chen, Z.; Zang, H.; Fan, S.; Gu, J.; Zhang, G.; Sun, K.D.; Wang, Q.; He, Y.; Owonikoko, T.K.; et al. Targeting c-Myc to Overcome Acquired Resistance of EGFR Mutant NSCLC Cells to the Third-Generation EGFR Tyrosine Kinase Inhibitor, Osimertinib. Cancer Res. 2021, 81, 4822–4834. [Google Scholar] [CrossRef] [PubMed]
  66. Jonna, S.; Subramaniam, D.S. Molecular diagnostics and targeted therapies in non-small cell lung cancer (NSCLC): An update. Discov. Med. 2019, 27, 167–170. [Google Scholar] [PubMed]
  67. Lei, X.; Li, Y.F.; Chen, G.D.; Ou, D.P.; Qiu, X.X.; Zuo, C.H.; Yang, L.Y. Ack1 overexpression promotes metastasis and indicates poor prognosis of hepatocellular carcinoma. Oncotarget 2015, 6, 40622–40641. [Google Scholar] [CrossRef] [PubMed]
  68. Karaca, M.; Liu, Y.; Zhang, Z.; De Silva, D.; Parker, J.S.; Earp, H.S.; Whang, Y.E. Mutation of androgen receptor N-terminal phosphorylation site Tyr-267 leads to inhibition of nuclear translocation and DNA binding. PLoS ONE 2015, 10, e0126270. [Google Scholar] [CrossRef] [PubMed]
  69. Su, B.; Gillard, B.; Gao, L.; Eng, K.H.; Gelman, I.H. Src controls castration recurrence of CWR22 prostate cancer xenografts. Cancer Med. 2013, 2, 784–792. [Google Scholar] [CrossRef]
  70. Hu, X.; Li, J.; Fu, M.; Zhao, X.; Wang, W. The JAK/STAT signaling pathway: From bench to clinic. Signal Transduct. Target. Ther. 2021, 6, 402. [Google Scholar] [CrossRef]
  71. Ungureanu, D.; Wu, J.; Pekkala, T.; Niranjan, Y.; Young, C.; Jensen, O.N.; Xu, C.F.; Neubert, T.A.; Skoda, R.C.; Hubbard, S.R.; et al. The pseudokinase domain of JAK2 is a dual-specificity protein kinase that negatively regulates cytokine signaling. Nat. Struct. Mol. Biol. 2011, 18, 971–976. [Google Scholar] [CrossRef] [PubMed]
  72. Saharinen, P.; Silvennoinen, O. The pseudokinase domain is required for suppression of basal activity of Jak2 and Jak3 tyrosine kinases and for cytokine-inducible activation of signal transduction. J. Biol. Chem. 2002, 277, 47954–47963. [Google Scholar] [CrossRef] [PubMed]
  73. Glassman, C.R.; Tsutsumi, N.; Saxton, R.A.; Lupardus, P.J.; Jude, K.M.; Garcia, K.C. Structure of a Janus kinase cytokine receptor complex reveals the basis for dimeric activation. Science 2022, 376, 163–169. [Google Scholar] [CrossRef] [PubMed]
  74. Haan, C.; Is’harc, H.; Hermanns, H.M.; Schmitz-Van De Leur, H.; Kerr, I.M.; Heinrich, P.C.; Grötzinger, J.; Behrmann, I. Mapping of a region within the N terminus of Jak1 involved in cytokine receptor interaction. J. Biol. Chem. 2001, 276, 37451–37458. [Google Scholar] [CrossRef] [PubMed]
  75. Giordanetto, F.; Kroemer, R.T. Prediction of the structure of human Janus kinase 2 (JAK2) comprising JAK homology domains 1 through 7. Protein Eng. 2002, 15, 727–737. [Google Scholar] [CrossRef] [PubMed]
  76. Ferrao, R.; Lupardus, P.J. The Janus Kinase (JAK) FERM and SH2 Domains: Bringing Specificity to JAK-Receptor Interactions. Front. Endocrinol. 2017, 8, 71. [Google Scholar] [CrossRef] [PubMed]
  77. Yunianto, I.; Currie, M.; Chitcholtan, K.; Sykes, P. Potential drug repurposing of ruxolitinib to inhibit the JAK/STAT pathway for the treatment of patients with epithelial ovarian cancer. J. Obstet. Gynaecol. Res. 2023, 49, 2563–2574. [Google Scholar] [CrossRef]
  78. Sheikh, E.; Tran, T.; Vranic, S.; Levy, A.; Bonfil, R.D. Role and significance of c-KIT receptor tyrosine kinase in cancer: A review. Bosn. J. Basic Med. Sci. 2022, 22, 683–698. [Google Scholar] [CrossRef]
  79. Ebrahimi, N.; Fardi, E.; Ghaderi, H.; Palizdar, S.; Khorram, R.; Vafadar, R.; Ghanaatian, M.; Rezaei-Tazangi, F.; Baziyar, P.; Ahmadi, A.; et al. Receptor tyrosine kinase inhibitors in cancer. Cell. Mol. Life Sci. 2023, 80, 104. [Google Scholar] [CrossRef]
  80. Pathania, S.; Pentikäinen, O.T.; Singh, P.K. A holistic view on c-Kit in cancer: Structure, signaling, pathophysiology and its inhibitors. Biochim. Biophys. Acta Rev. Cancer 2021, 1876, 188631. [Google Scholar] [CrossRef]
  81. Lin, J.X.; Leonard, W.J. The Common Cytokine Receptor γ Chain Family of Cytokines. Cold Spring Harb. Perspect. Biol. 2018, 10, a028449. [Google Scholar] [CrossRef] [PubMed]
  82. Giese, B.; Au-Yeung, C.K.; Herrmann, A.; Diefenbach, S.; Haan, C.; Küster, A.; Wortmann, S.B.; Roderburg, C.; Heinrich, P.C.; Behrmann, I.; et al. Long term association of the cytokine receptor gp130 and the Janus kinase Jak1 revealed by FRAP analysis. J. Biol. Chem. 2003, 278, 39205–39213. [Google Scholar] [CrossRef]
  83. Lensing, M.; Jabbari, A. An overview of JAK/STAT pathways and JAK inhibition in alopecia areata. Front. Immunol. 2022, 13, 955035. [Google Scholar] [CrossRef] [PubMed]
  84. Morris, R.; Kershaw, N.J.; Babon, J.J. The molecular details of cytokine signaling via the JAK/STAT pathway. Protein Sci. 2018, 27, 1984–2009. [Google Scholar] [CrossRef]
  85. Seif, F.; Khoshmirsafa, M.; Aazami, H.; Mohsenzadegan, M.; Sedighi, G.; Bahar, M. The role of JAK-STAT signaling pathway and its regulators in the fate of T helper cells. Cell Commun. Signal. 2017, 15, 23. [Google Scholar] [CrossRef]
  86. Kawamura, M.; McVicar, D.W.; Johnston, J.A.; Blake, T.B.; Chen, Y.Q.; Lal, B.K.; Lloyd, A.R.; Kelvin, D.J.; Staples, J.E.; Ortaldo, J.R. Molecular cloning of L-JAK, a Janus family protein-tyrosine kinase expressed in natural killer cells and activated leukocytes. Proc. Natl. Acad. Sci. USA 1994, 91, 6374–6378. [Google Scholar] [CrossRef]
  87. Musso, T.; Johnston, J.A.; Linnekin, D.; Varesio, L.; Rowe, T.K.; O’Shea, J.J.; McVicar, D.W. Regulation of JAK3 expression in human monocytes: Phosphorylation in response to interleukins 2, 4, and 7. J. Exp. Med. 1995, 181, 1425–1431. [Google Scholar] [CrossRef] [PubMed]
  88. Tortolani, P.J.; Lal, B.K.; Riva, A.; Johnston, J.A.; Chen, Y.Q.; Reaman, G.H.; Beckwith, M.; Longo, D.; Ortaldo, J.R.; Bhatia, K.; et al. Regulation of JAK3 expression and activation in human B cells and B cell malignancies. J. Immunol. 1995, 155, 5220–5226. [Google Scholar] [CrossRef]
  89. Hofmann, S.R.; Lam, A.Q.; Frank, S.; Zhou, Y.J.; Ramos, H.L.; Kanno, Y.; Agnello, D.; Youle, R.J.; O’Shea, J.J. Jak3-independent trafficking of the common gamma chain receptor subunit: Chaperone function of Jaks revisited. Mol. Cell. Biol. 2004, 24, 5039–5049. [Google Scholar] [CrossRef]
  90. David, M.; Zhou, G.; Pine, R.; Dixon, J.E.; Larner, A.C. The SH2 domain-containing tyrosine phosphatase PTP1D is required for interferon alpha/beta-induced gene expression. J. Biol. Chem. 1996, 271, 15862–15865. [Google Scholar] [CrossRef]
  91. Abramovich, C.; Yakobson, B.; Chebath, J.; Revel, M. A protein-arginine methyltransferase binds to the intracytoplasmic domain of the IFNAR1 chain in the type I interferon receptor. EMBO J. 1997, 16, 260–266. [Google Scholar] [CrossRef] [PubMed]
  92. Silvennoinen, O.; Ihle, J.N.; Schlessinger, J.; Levy, D.E. Interferon-induced nuclear signalling by Jak protein tyrosine kinases. Nature 1993, 366, 583–585. [Google Scholar] [CrossRef] [PubMed]
  93. Ge, T.; Jhala, G.; Fynch, S.; Akazawa, S.; Litwak, S.; Pappas, E.G.; Catterall, T.; Vakil, I.; Long, A.J.; Olson, L.M.; et al. The JAK1 Selective Inhibitor ABT 317 Blocks Signaling Through Interferon-γ and Common γ Chain Cytokine Receptors to Reverse Autoimmune Diabetes in NOD Mice. Front. Immunol. 2020, 11, 588543. [Google Scholar] [CrossRef] [PubMed]
  94. Leonard, W.J.; Lin, J.X.; O’Shea, J.J. The γ. Immunity 2019, 50, 832–850. [Google Scholar] [CrossRef]
  95. Waickman, A.T.; Park, J.Y.; Park, J.H. The common γ-chain cytokine receptor: Tricks-and-treats for T cells. Cell. Mol. Life Sci. 2016, 73, 253–269. [Google Scholar] [CrossRef] [PubMed]
  96. Rose, T.M.; Bruce, A.G. Oncostatin M is a member of a cytokine family that includes leukemia-inhibitory factor, granulocyte colony-stimulating factor, and interleukin 6. Proc. Natl. Acad. Sci. USA 1991, 88, 8641–8645. [Google Scholar] [CrossRef]
  97. Waters, M.J.; Brooks, A.J. JAK2 activation by growth hormone and other cytokines. Biochem. J. 2015, 466, 1–11. [Google Scholar] [CrossRef] [PubMed]
  98. Prchal-Murphy, M.; Semper, C.; Lassnig, C.; Wallner, B.; Gausterer, C.; Teppner-Klymiuk, I.; Kobolak, J.; Müller, S.; Kolbe, T.; Karaghiosoff, M.; et al. TYK2 kinase activity is required for functional type I interferon responses in vivo. PLoS ONE 2012, 7, e39141. [Google Scholar] [CrossRef]
  99. Ma, C.S.; Avery, D.T.; Chan, A.; Batten, M.; Bustamante, J.; Boisson-Dupuis, S.; Arkwright, P.D.; Kreins, A.Y.; Averbuch, D.; Engelhard, D.; et al. Functional STAT3 deficiency compromises the generation of human T follicular helper cells. Blood 2012, 119, 3997–4008. [Google Scholar] [CrossRef]
  100. Gracey, E.; Hromadová, D.; Lim, M.; Qaiyum, Z.; Zeng, M.; Yao, Y.; Srinath, A.; Baglaenko, Y.; Yeremenko, N.; Westlin, W.; et al. TYK2 inhibition reduces type 3 immunity and modifies disease progression in murine spondyloarthritis. J. Clin. Investig. 2020, 130, 1863–1878. [Google Scholar] [CrossRef]
  101. Järvinen, T.M.; Hellquist, A.; Koskenmies, S.; Einarsdottir, E.; Koskinen, L.L.; Jeskanen, L.; Berglind, L.; Panelius, J.; Hasan, T.; Ranki, A.; et al. Tyrosine kinase 2 and interferon regulatory factor 5 polymorphisms are associated with discoid and subacute cutaneous lupus erythematosus. Exp. Dermatol. 2010, 19, 123–131. [Google Scholar] [CrossRef] [PubMed]
  102. Jehanno, C.; Vulin, M.; Richina, V.; Richina, F.; Bentires-Alj, M. Phenotypic plasticity during metastatic colonization. Trends Cell Biol. 2022, 32, 854–867. [Google Scholar] [CrossRef] [PubMed]
  103. Westneat, D.F.; Potts, L.J.; Sasser, K.L.; Shaffer, J.D. Causes and Consequences of Phenotypic Plasticity in Complex Environments. Trends Ecol. Evol. 2019, 34, 555–568. [Google Scholar] [CrossRef] [PubMed]
  104. Babon, J.J.; Lucet, I.S.; Murphy, J.M.; Nicola, N.A.; Varghese, L.N. The molecular regulation of Janus kinase (JAK) activation. Biochem. J. 2014, 462, 1–13. [Google Scholar] [CrossRef] [PubMed]
  105. Liau, N.P.D.; Laktyushin, A.; Morris, R.; Sandow, J.J.; Nicola, N.A.; Kershaw, N.J.; Babon, J.J. Enzymatic Characterization of Wild-Type and Mutant Janus Kinase 1. Cancers 2019, 11, 1701. [Google Scholar] [CrossRef] [PubMed]
  106. Muromoto, R.; Oritani, K.; Matsuda, T. Current understanding of the role of tyrosine kinase 2 signaling in immune responses. World J. Biol. Chem. 2022, 13, 1–14. [Google Scholar] [CrossRef] [PubMed]
  107. Fortelny, N.; Farlik, M.; Fife, V.; Gorki, A.D.; Lassnig, C.; Maurer, B.; Meissl, K.; Dolezal, M.; Boccuni, L.; Ravi Sundar Jose Geetha, A.; et al. JAK-STAT signaling maintains homeostasis in T cells and macrophages. Nat. Immunol. 2024, 25, 847–859. [Google Scholar] [CrossRef] [PubMed]
  108. Stanhope, B.A.; Jaggard, J.B.; Gratton, M.; Brown, E.B.; Keene, A.C. Sleep Regulates Glial Plasticity and Expression of the Engulfment Receptor Draper Following Neural Injury. Curr. Biol. 2020, 30, 1092–1101.e1093. [Google Scholar] [CrossRef] [PubMed]
  109. Groffen, J.; Heisterkamp, N.; Shibuya, M.; Hanafusa, H.; Stephenson, J.R. Transforming genes of avian (v-fps) and mammalian (v-fes) retroviruses correspond to a common cellular locus. Virology 1983, 125, 480–486. [Google Scholar] [CrossRef]
  110. Craig, A.W. FES/FER kinase signaling in hematopoietic cells and leukemias. Front. Biosci. 2012, 17, 861–875. [Google Scholar] [CrossRef]
  111. Laurent, C.E.; Delfino, F.J.; Cheng, H.Y.; Smithgall, T.E. The human c-Fes tyrosine kinase binds tubulin and microtubules through separate domains and promotes microtubule assembly. Mol. Cell. Biol. 2004, 24, 9351–9358. [Google Scholar] [CrossRef] [PubMed]
  112. Hellwig, S.; Miduturu, C.V.; Kanda, S.; Zhang, J.; Filippakopoulos, P.; Salah, E.; Deng, X.; Choi, H.G.; Zhou, W.; Hur, W.; et al. Small-molecule inhibitors of the c-Fes protein-tyrosine kinase. Chem. Biol. 2012, 19, 529–540. [Google Scholar] [CrossRef] [PubMed]
  113. Filippakopoulos, P.; Kofler, M.; Hantschel, O.; Gish, G.D.; Grebien, F.; Salah, E.; Neudecker, P.; Kay, L.E.; Turk, B.E.; Superti-Furga, G.; et al. Structural coupling of SH2-kinase domains links Fes and Abl substrate recognition and kinase activation. Cell 2008, 134, 793–803. [Google Scholar] [CrossRef] [PubMed]
  114. Rogers, J.A.; Read, R.D.; Li, J.; Peters, K.L.; Smithgall, T.E. Autophosphorylation of the Fes tyrosine kinase. Evidence for an intermolecular mechanism involving two kinase domain tyrosine residues. J. Biol. Chem. 1996, 271, 17519–17525. [Google Scholar] [CrossRef]
  115. Menegon, A.; Burgaya, F.; Baudot, P.; Dunlap, D.D.; Girault, J.A.; Valtorta, F. FAK+ and PYK2/CAKbeta, two related tyrosine kinases highly expressed in the central nervous system: Similarities and differences in the expression pattern. Eur. J. Neurosci. 1999, 11, 3777–3788. [Google Scholar] [CrossRef] [PubMed]
  116. Naser, R.; Aldehaiman, A.; Díaz-Galicia, E.; Arold, S.T. Endogenous Control Mechanisms of FAK and PYK2 and Their Relevance to Cancer Development. Cancers 2018, 10, 196. [Google Scholar] [CrossRef] [PubMed]
  117. Dunty, J.M.; Schaller, M.D. The N termini of focal adhesion kinase family members regulate substrate phosphorylation, localization, and cell morphology. J. Biol. Chem. 2002, 277, 45644–45654. [Google Scholar] [CrossRef] [PubMed]
  118. Palmer, R.H.; Fessler, L.I.; Edeen, P.T.; Madigan, S.J.; McKeown, M.; Hunter, T. DFak56 is a novel Drosophila melanogaster focal adhesion kinase. J. Biol. Chem. 1999, 274, 35621–35629. [Google Scholar] [CrossRef] [PubMed]
  119. Cooley, M.A.; Broome, J.M.; Ohngemach, C.; Romer, L.H.; Schaller, M.D. Paxillin binding is not the sole determinant of focal adhesion localization or dominant-negative activity of focal adhesion kinase/focal adhesion kinase-related nonkinase. Mol. Biol. Cell 2000, 11, 3247–3263. [Google Scholar] [CrossRef]
  120. Bolós, V.; Gasent, J.M.; López-Tarruella, S.; Grande, E. The dual kinase complex FAK-Src as a promising therapeutic target in cancer. OncoTargets Ther. 2010, 3, 83–97. [Google Scholar] [CrossRef]
  121. Borkúti, P.; Kristó, I.; Szabó, A.; Kovács, Z.; Vilmos, P. FERM domain-containing proteins are active components of the cell nucleus. Life Sci. Alliance 2024, 7, e202302489. [Google Scholar] [CrossRef] [PubMed]
  122. Pearson, M.A.; Reczek, D.; Bretscher, A.; Karplus, P.A. Structure of the ERM protein moesin reveals the FERM domain fold masked by an extended actin binding tail domain. Cell 2000, 101, 259–270. [Google Scholar] [CrossRef] [PubMed]
  123. Loving, H.S.; Underbakke, E.S. Conformational Dynamics of FERM-Mediated Autoinhibition in Pyk2 Tyrosine Kinase. Biochemistry 2019, 58, 3767–3776. [Google Scholar] [CrossRef] [PubMed]
  124. Tapial Martínez, P.; López Navajas, P.; Lietha, D. FAK Structure and Regulation by Membrane Interactions and Force in Focal Adhesions. Biomolecules 2020, 10, 179. [Google Scholar] [CrossRef]
  125. Prutzman, K.C.; Gao, G.; King, M.L.; Iyer, V.V.; Mueller, G.A.; Schaller, M.D.; Campbell, S.L. The focal adhesion targeting domain of focal adhesion kinase contains a hinge region that modulates tyrosine 926 phosphorylation. Structure 2004, 12, 881–891. [Google Scholar] [CrossRef] [PubMed]
  126. Shen, Y.; Schaller, M.D. Focal adhesion targeting: The critical determinant of FAK regulation and substrate phosphorylation. Mol. Biol. Cell 1999, 10, 2507–2518. [Google Scholar] [CrossRef] [PubMed]
  127. Kadaré, G.; Gervasi, N.; Brami-Cherrier, K.; Blockus, H.; El Messari, S.; Arold, S.T.; Girault, J.A. Conformational dynamics of the focal adhesion targeting domain control specific functions of focal adhesion kinase in cells. J. Biol. Chem. 2015, 290, 478–491. [Google Scholar] [CrossRef] [PubMed]
  128. Subauste, M.C.; Pertz, O.; Adamson, E.D.; Turner, C.E.; Junger, S.; Hahn, K.M. Vinculin modulation of paxillin-FAK interactions regulates ERK to control survival and motility. J. Cell Biol. 2004, 165, 371–381. [Google Scholar] [CrossRef] [PubMed]
  129. Deramaudt, T.B.; Dujardin, D.; Noulet, F.; Martin, S.; Vauchelles, R.; Takeda, K.; Rondé, P. Altering FAK-paxillin interactions reduces adhesion, migration and invasion processes. PLoS ONE 2014, 9, e92059. [Google Scholar] [CrossRef]
  130. Thomas, J.W.; Cooley, M.A.; Broome, J.M.; Salgia, R.; Griffin, J.D.; Lombardo, C.R.; Schaller, M.D. The role of focal adhesion kinase binding in the regulation of tyrosine phosphorylation of paxillin. J. Biol. Chem. 1999, 274, 36684–36692. [Google Scholar] [CrossRef]
  131. Birge, R.B.; Kalodimos, C.; Inagaki, F.; Tanaka, S. Crk and CrkL adaptor proteins: Networks for physiological and pathological signaling. Cell Commun. Signal 2009, 7, 13. [Google Scholar] [CrossRef]
  132. Schlaepfer, D.D.; Hauck, C.R.; Sieg, D.J. Signaling through focal adhesion kinase. Prog. Biophys. Mol. Biol. 1999, 71, 435–478. [Google Scholar] [CrossRef] [PubMed]
  133. Rigiracciolo, D.C.; Cirillo, F.; Talia, M.; Muglia, L.; Gutkind, J.S.; Maggiolini, M.; Lappano, R. Focal Adhesion Kinase Fine Tunes Multifaced Signals toward Breast Cancer Progression. Cancers 2021, 13, 645. [Google Scholar] [CrossRef] [PubMed]
  134. Golubovskaya, V.M. Targeting FAK in human cancer: From finding to first clinical trials. Front. Biosci. 2014, 19, 687–706. [Google Scholar] [CrossRef] [PubMed]
  135. Pylayeva, Y.; Gillen, K.M.; Gerald, W.; Beggs, H.E.; Reichardt, L.F.; Giancotti, F.G. Ras- and PI3K-dependent breast tumorigenesis in mice and humans requires focal adhesion kinase signaling. J. Clin. Investig. 2009, 119, 252–266. [Google Scholar] [CrossRef] [PubMed]
  136. Chuang, H.H.; Zhen, Y.Y.; Tsai, Y.C.; Chuang, C.H.; Hsiao, M.; Huang, M.S.; Yang, C.J. FAK in Cancer: From Mechanisms to Therapeutic Strategies. Int. J. Mol. Sci. 2022, 23, 1726. [Google Scholar] [CrossRef] [PubMed]
  137. Mano, H. Tec family of protein-tyrosine kinases: An overview of their structure and function. Cytokine Growth Factor Rev. 1999, 10, 267–280. [Google Scholar] [CrossRef] [PubMed]
  138. Hyvönen, M.; Saraste, M. Structure of the PH domain and Btk motif from Bruton’s tyrosine kinase: Molecular explanations for X-linked agammaglobulinaemia. EMBO J. 1997, 16, 3396–3404. [Google Scholar] [CrossRef] [PubMed]
  139. Lin, D.Y.; Kueffer, L.E.; Juneja, P.; Wales, T.E.; Engen, J.R.; Andreotti, A.H. Conformational heterogeneity of the BTK PHTH domain drives multiple regulatory states. eLife 2024, 12, RP89489. [Google Scholar] [CrossRef]
  140. Roberts, J.M.; Tarafdar, S.; Joseph, R.E.; Andreotti, A.H.; Smithgall, T.E.; Engen, J.R.; Wales, T.E. Dynamics of the Tec-family tyrosine kinase SH3 domains. Protein Sci. 2016, 25, 852–864. [Google Scholar] [CrossRef]
  141. Joseph, R.E.; Andreotti, A.H. Conformational snapshots of Tec kinases during signaling. Immunol. Rev. 2009, 228, 74–92. [Google Scholar] [CrossRef] [PubMed]
  142. Yoshida, K.; Yamashita, Y.; Miyazato, A.; Ohya, K.; Kitanaka, A.; Ikeda, U.; Shimada, K.; Yamanaka, T.; Ozawa, K.; Mano, H. Mediation by the protein-tyrosine kinase Tec of signaling between the B cell antigen receptor and Dok-1. J. Biol. Chem. 2000, 275, 24945–24952. [Google Scholar] [CrossRef] [PubMed]
  143. Ortutay, C.; Nore, B.F.; Vihinen, M.; Smith, C.I. Phylogeny of Tec family kinases identification of a premetazoan origin of Btk, Bmx, Itk, Tec, Txk, and the Btk regulator SH3BP5. Adv. Genet. 2008, 64, 51–80. [Google Scholar] [CrossRef] [PubMed]
  144. Andersen, T.C.B.; Kristiansen, P.E.; Huszenicza, Z.; Johansson, M.U.; Gopalakrishnan, R.P.; Kjelstrup, H.; Boyken, S.; Sundvold-Gjerstad, V.; Granum, S.; Sørli, M.; et al. The SH3 domains of the protein kinases ITK and LCK compete for adjacent sites on T cell-specific adapter protein. J. Biol. Chem. 2019, 294, 15480–15494. [Google Scholar] [CrossRef] [PubMed]
  145. Kashiwakura, J.; Suzuki, N.; Nagafuchi, H.; Takeno, M.; Takeba, Y.; Shimoyama, Y.; Sakane, T. Txk, a nonreceptor tyrosine kinase of the Tec family, is expressed in T helper type 1 cells and regulates interferon gamma production in human T lymphocytes. J. Exp. Med. 1999, 190, 1147–1154. [Google Scholar] [CrossRef] [PubMed]
  146. Qiu, Y.; Kung, H.J. Signaling network of the Btk family kinases. Oncogene 2000, 19, 5651–5661. [Google Scholar] [CrossRef] [PubMed]
  147. Eldeeb, M.A.; Fahlman, R.P. Phosphorylation Impacts N-end Rule Degradation of the Proteolytically Activated Form of BMX Kinase. J. Biol. Chem. 2016, 291, 22757–22768. [Google Scholar] [CrossRef] [PubMed]
  148. Bao, Y.; Zheng, J.; Han, C.; Jin, J.; Han, H.; Liu, Y.; Lau, Y.L.; Tu, W.; Cao, X. Tyrosine kinase Btk is required for NK cell activation. J. Biol. Chem. 2012, 287, 23769–23778. [Google Scholar] [CrossRef] [PubMed]
  149. Brunner, C.; Müller, B.; Wirth, T. Bruton’s Tyrosine Kinase is involved in innate and adaptive immunity. Histol. Histopathol. 2005, 20, 945–955. [Google Scholar] [CrossRef]
  150. Arcaro, A.; Guerreiro, A.S. The phosphoinositide 3-kinase pathway in human cancer: Genetic alterations and therapeutic implications. Curr. Genom. 2007, 8, 271–306. [Google Scholar] [CrossRef]
  151. Yang, W.C.; Ching, K.A.; Tsoukas, C.D.; Berg, L.J. Tec kinase signaling in T cells is regulated by phosphatidylinositol 3-kinase and the Tec pleckstrin homology domain. J. Immunol. 2001, 166, 387–395. [Google Scholar] [CrossRef] [PubMed]
  152. Boggon, T.J.; Eck, M.J. Structure and regulation of Src family kinases. Oncogene 2004, 23, 7918–7927. [Google Scholar] [CrossRef] [PubMed]
  153. Engen, J.R.; Wales, T.E.; Hochrein, J.M.; Meyn, M.A.; Banu Ozkan, S.; Bahar, I.; Smithgall, T.E. Structure and dynamic regulation of Src-family kinases. Cell. Mol. Life Sci. 2008, 65, 3058–3073. [Google Scholar] [CrossRef] [PubMed]
  154. Ortiz, M.A.; Mikhailova, T.; Li, X.; Porter, B.A.; Bah, A.; Kotula, L. Src family kinases, adaptor proteins and the actin cytoskeleton in epithelial-to-mesenchymal transition. Cell Commun. Signal. 2021, 19, 67. [Google Scholar] [CrossRef]
  155. Amata, I.; Maffei, M.; Pons, M. Phosphorylation of unique domains of Src family kinases. Front. Genet. 2014, 5, 181. [Google Scholar] [CrossRef] [PubMed]
  156. Voisset, E.; Brenet, F.; Lopez, S.; de Sepulveda, P. SRC-Family Kinases in Acute Myeloid Leukaemia and Mastocytosis. Cancers 2020, 12, 1996. [Google Scholar] [CrossRef] [PubMed]
  157. Okada, M. Regulation of the SRC family kinases by Csk. Int. J. Biol. Sci. 2012, 8, 1385–1397. [Google Scholar] [CrossRef] [PubMed]
  158. Kovács, M.; Németh, T.; Jakus, Z.; Sitaru, C.; Simon, E.; Futosi, K.; Botz, B.; Helyes, Z.; Lowell, C.A.; Mócsai, A. The Src family kinases Hck, Fgr, and Lyn are critical for the generation of the in vivo inflammatory environment without a direct role in leukocyte recruitment. J. Exp. Med. 2014, 211, 1993–2011. [Google Scholar] [CrossRef] [PubMed]
  159. Marhäll, A.; Kazi, J.U.; Rönnstrand, L. The Src family kinase LCK cooperates with oncogenic FLT3/ITD in cellular transformation. Sci. Rep. 2017, 7, 13734. [Google Scholar] [CrossRef]
  160. Pestina, T.I.; Stenberg, P.E.; Druker, B.J.; Steward, S.A.; Hutson, N.K.; Barrie, R.J.; Jackson, C.W. Identification of the Src family kinases, Lck and Fgr in platelets. Their tyrosine phosphorylation status and subcellular distribution compared with other Src family members. Arterioscler. Thromb. Vasc. Biol. 1997, 17, 3278–3285. [Google Scholar] [CrossRef]
  161. Grgurevich, S.; Linnekin, D.; Musso, T.; Zhang, X.; Modi, W.; Varesio, L.; Ruscetti, F.W.; Ortaldo, J.R.; McVicar, D.W. The Csk-like proteins Lsk, Hyl, and Matk represent the same Csk homologous kinase (Chk) and are regulated by stem cell factor in the megakaryoblastic cell line MO7e. Growth Factors 1997, 14, 103–115. [Google Scholar] [CrossRef]
  162. Avraham, S.; Jiang, S.; Ota, S.; Fu, Y.; Deng, B.; Dowler, L.L.; White, R.A.; Avraham, H. Structural and functional studies of the intracellular tyrosine kinase MATK gene and its translated product. J. Biol. Chem. 1995, 270, 1833–1842. [Google Scholar] [CrossRef]
  163. Stenberg, P.E.; Pestina, T.I.; Barrie, R.J.; Jackson, C.W. The Src family kinases, Fgr, Fyn, Lck, and Lyn, colocalize with coated membranes in platelets. Blood 1997, 89, 2384–2393. [Google Scholar] [CrossRef]
  164. Advani, G.; Lim, Y.C.; Catimel, B.; Lio, D.S.S.; Ng, N.L.Y.; Chüeh, A.C.; Tran, M.; Anasir, M.I.; Verkade, H.; Zhu, H.J.; et al. Csk-homologous kinase (Chk) is an efficient inhibitor of Src-family kinases but a poor catalyst of phosphorylation of their C-terminal regulatory tyrosine. Cell Commun. Signal. 2017, 15, 29. [Google Scholar] [CrossRef] [PubMed]
  165. Shah, N.H.; Amacher, J.F.; Nocka, L.M.; Kuriyan, J. The Src module: An ancient scaffold in the evolution of cytoplasmic tyrosine kinases. Crit. Rev. Biochem. Mol. Biol. 2018, 53, 535–563. [Google Scholar] [CrossRef] [PubMed]
  166. Goel, R.K.; Lukong, K.E. Tracing the footprints of the breast cancer oncogene BRK-Past till present. Biochim. Biophys. Acta 2015, 1856, 39–54. [Google Scholar] [CrossRef]
  167. Goel, R.K.; Kim, N.; Lukong, K.E. Seeking a better understanding of the non-receptor tyrosine kinase, SRMS. Heliyon 2023, 9, e16421. [Google Scholar] [CrossRef] [PubMed]
  168. Bagnato, G.; Leopizzi, M.; Urciuoli, E.; Peruzzi, B. Nuclear Functions of the Tyrosine Kinase Src. Int. J. Mol. Sci. 2020, 21, 2675. [Google Scholar] [CrossRef]
  169. Goel, R.K.; Miah, S.; Black, K.; Kalra, N.; Dai, C.; Lukong, K.E. The unique N-terminal region of SRMS regulates enzymatic activity and phosphorylation of its novel substrate docking protein 1. FEBS J. 2013, 280, 4539–4559. [Google Scholar] [CrossRef]
  170. Zhai, Y.; Yang, J.; Zhang, J.; Li, Q.; Zheng, T. Src-family Protein Tyrosine Kinases: A promising target for treating Cardiovascular Diseases. Int. J. Med. Sci. 2021, 18, 1216–1224. [Google Scholar] [CrossRef]
  171. Sen, B.; Johnson, F.M. Regulation of SRC family kinases in human cancers. J. Signal Transduct. 2011, 2011, 865819. [Google Scholar] [CrossRef] [PubMed]
  172. Kinoshita-Kikuta, E.; Utsumi, T.; Miyazaki, A.; Tokumoto, C.; Doi, K.; Harada, H.; Kinoshita, E.; Koike, T. Protein-N-myristoylation-dependent phosphorylation of serine 13 of tyrosine kinase Lyn by casein kinase 1γ at the Golgi during intracellular protein traffic. Sci. Rep. 2020, 10, 16273. [Google Scholar] [CrossRef] [PubMed]
  173. Berclaz, G.; Altermatt, H.J.; Rohrbach, V.; Dreher, E.; Ziemiecki, A.; Andres, A.C. Hormone-dependent nuclear localization of the tyrosine kinase iyk in the normal human breast epithelium and loss of expression during carcinogenesis. Int. J. Cancer 2000, 85, 889–894. [Google Scholar] [CrossRef]
  174. McClendon, C.J.; Miller, W.T. Structure, Function, and Regulation of the SRMS Tyrosine Kinase. Int. J. Mol. Sci. 2020, 21, 4233. [Google Scholar] [CrossRef] [PubMed]
  175. Liang, X.; Lu, Y.; Wilkes, M.; Neubert, T.A.; Resh, M.D. The N-terminal SH4 region of the Src family kinase Fyn is modified by methylation and heterogeneous fatty acylation: Role in membrane targeting, cell adhesion, and spreading. J. Biol. Chem. 2004, 279, 8133–8139. [Google Scholar] [CrossRef] [PubMed]
  176. Fhu, C.W.; Ali, A. Protein Lipidation by Palmitoylation and Myristoylation in Cancer. Front. Cell Dev. Biol. 2021, 9, 673647. [Google Scholar] [CrossRef] [PubMed]
  177. Annerén, C.; Welsh, M.; Jansson, L. Glucose intolerance and reduced islet blood flow in transgenic mice expressing the FRK tyrosine kinase under the control of the rat insulin promoter. Am. J. Physiol. Endocrinol. Metab. 2007, 292, E1183–E1190. [Google Scholar] [CrossRef]
  178. Gu, J.J.; Ryu, J.R.; Pendergast, A.M. Abl tyrosine kinases in T-cell signaling. Immunol. Rev. 2009, 228, 170–183. [Google Scholar] [CrossRef]
  179. Greuber, E.K.; Pendergast, A.M. Abl family kinases regulate FcγR-mediated phagocytosis in murine macrophages. J. Immunol. 2012, 189, 5382–5392. [Google Scholar] [CrossRef]
  180. Madan, S.; Kron, B.; Jin, Z.; Al Shamy, G.; Campeau, P.M.; Sun, Q.; Chen, S.; Cherian, L.; Chen, Y.; Munivez, E.; et al. Arginase overexpression in neurons and its effect on traumatic brain injury. Mol. Genet. Metab. 2018, 125, 112–117. [Google Scholar] [CrossRef]
  181. Stratoulias, V.; Ruiz, R.; Kanatani, S.; Osman, A.M.; Keane, L.; Armengol, J.A.; Rodríguez-Moreno, A.; Murgoci, A.N.; García-Domínguez, I.; Alonso-Bellido, I.; et al. ARG1-expressing microglia show a distinct molecular signature and modulate postnatal development and function of the mouse brain. Nat. Neurosci. 2023, 26, 1008–1020. [Google Scholar] [CrossRef] [PubMed]
  182. Lorenz, S.; Deng, P.; Hantschel, O.; Superti-Furga, G.; Kuriyan, J. Crystal structure of an SH2-kinase construct of c-Abl and effect of the SH2 domain on kinase activity. Biochem. J. 2015, 468, 283–291. [Google Scholar] [CrossRef] [PubMed]
  183. Panjarian, S.; Iacob, R.E.; Chen, S.; Engen, J.R.; Smithgall, T.E. Structure and dynamic regulation of Abl kinases. J. Biol. Chem. 2013, 288, 5443–5450. [Google Scholar] [CrossRef] [PubMed]
  184. Preyer, M.; Vigneri, P.; Wang, J.Y. Interplay between kinase domain autophosphorylation and F-actin binding domain in regulating imatinib sensitivity and nuclear import of BCR-ABL. PLoS ONE 2011, 6, e17020. [Google Scholar] [CrossRef] [PubMed]
  185. Colicelli, J. ABL tyrosine kinases: Evolution of function, regulation, and specificity. Sci. Signal. 2010, 3, re6. [Google Scholar] [CrossRef] [PubMed]
  186. Underhill-Day, N.; Pierce, A.; Thompson, S.E.; Xenaki, D.; Whetton, A.D.; Owen-Lynch, P.J. Role of the C-terminal actin binding domain in BCR/ABL-mediated survival and drug resistance. Br. J. Haematol. 2006, 132, 774–783. [Google Scholar] [CrossRef] [PubMed]
  187. Van Etten, R.A.; Jackson, P.K.; Baltimore, D.; Sanders, M.C.; Matsudaira, P.T.; Janmey, P.A. The COOH terminus of the c-Abl tyrosine kinase contains distinct F- and G-actin binding domains with bundling activity. J. Cell Biol. 1994, 124, 325–340. [Google Scholar] [CrossRef] [PubMed]
  188. Taagepera, S.; McDonald, D.; Loeb, J.E.; Whitaker, L.L.; McElroy, A.K.; Wang, J.Y.; Hope, T.J. Nuclear-cytoplasmic shuttling of C-ABL tyrosine kinase. Proc. Natl. Acad. Sci. USA 1998, 95, 7457–7462. [Google Scholar] [CrossRef]
  189. Cao, C.; Leng, Y.; Li, C.; Kufe, D. Functional interaction between the c-Abl and Arg protein-tyrosine kinases in the oxidative stress response. J. Biol. Chem. 2003, 278, 12961–12967. [Google Scholar] [CrossRef]
  190. Buffa, P.; Manzella, L.; Consoli, M.L.; Messina, A.; Vigneri, P. Modelling of the ABL and ARG proteins predicts two functionally critical regions that are natively unfolded. Proteins 2007, 67, 1–11. [Google Scholar] [CrossRef]
  191. Wang, B.; Golemis, E.A.; Kruh, G.D. ArgBP2, a multiple Src homology 3 domain-containing, Arg/Abl-interacting protein, is phosphorylated in v-Abl-transformed cells and localized in stress fibers and cardiocyte Z-disks. J. Biol. Chem. 1997, 272, 17542–17550. [Google Scholar] [CrossRef]
  192. Miao, Y.J.; Wang, J.Y. Binding of A/T-rich DNA by three high mobility group-like domains in c-Abl tyrosine kinase. J. Biol. Chem. 1996, 271, 22823–22830. [Google Scholar] [CrossRef]
  193. MacGrath, S.M.; Koleske, A.J. Arg/Abl2 modulates the affinity and stoichiometry of binding of cortactin to F-actin. Biochemistry 2012, 51, 6644–6653. [Google Scholar] [CrossRef] [PubMed]
  194. Kerrisk, M.E.; Koleske, A.J. Arg kinase signaling in dendrite and synapse stabilization pathways: Memory, cocaine sensitivity, and stress. Int. J. Biochem. Cell Biol. 2013, 45, 2496–2500. [Google Scholar] [CrossRef]
  195. Wetzler, M.; Talpaz, M.; Van Etten, R.A.; Hirsh-Ginsberg, C.; Beran, M.; Kurzrock, R. Subcellular localization of Bcr, Abl, and Bcr-Abl proteins in normal and leukemic cells and correlation of expression with myeloid differentiation. J. Clin. Investig. 1993, 92, 1925–1939. [Google Scholar] [CrossRef] [PubMed]
  196. Knauer, S.K.; Moodt, S.; Berg, T.; Liebel, U.; Pepperkok, R.; Stauber, R.H. Translocation biosensors to study signal-specific nucleo-cytoplasmic transport, protease activity and protein-protein interactions. Traffic 2005, 6, 594–606. [Google Scholar] [CrossRef]
  197. de Buhr, S.; Gräter, F. Myristoyl’s dual role in allosterically regulating and localizing Abl kinase. eLife 2023, 12, e85216. [Google Scholar] [CrossRef]
  198. Woodring, P.J.; Litwack, E.D.; O’Leary, D.D.; Lucero, G.R.; Wang, J.Y.; Hunter, T. Modulation of the F-actin cytoskeleton by c-Abl tyrosine kinase in cell spreading and neurite extension. J. Cell Biol. 2002, 156, 879–892. [Google Scholar] [CrossRef] [PubMed]
  199. Sato, M.; Maruoka, M.; Takeya, T. Functional mechanisms and roles of adaptor proteins in abl-regulated cytoskeletal actin dynamics. J. Signal Transduct. 2012, 2012, 414913. [Google Scholar] [CrossRef]
  200. Aoyama, K.; Yuki, R.; Horiike, Y.; Kubota, S.; Yamaguchi, N.; Morii, M.; Ishibashi, K.; Nakayama, Y.; Kuga, T.; Hashimoto, Y.; et al. Formation of long and winding nuclear F-actin bundles by nuclear c-Abl tyrosine kinase. Exp. Cell Res. 2013, 319, 3251–3268. [Google Scholar] [CrossRef]
  201. Hu, H.; Bliss, J.M.; Wang, Y.; Colicelli, J. RIN1 is an ABL tyrosine kinase activator and a regulator of epithelial-cell adhesion and migration. Curr. Biol. 2005, 15, 815–823. [Google Scholar] [CrossRef]
  202. Sriram, G.; Reichman, C.; Tunceroglu, A.; Kaushal, N.; Saleh, T.; Machida, K.; Mayer, B.; Ge, Q.; Li, J.; Hornbeck, P.; et al. Phosphorylation of Crk on tyrosine 251 in the RT loop of the SH3C domain promotes Abl kinase transactivation. Oncogene 2011, 30, 4645–4655. [Google Scholar] [CrossRef]
  203. Cao, X.; Tanis, K.Q.; Koleske, A.J.; Colicelli, J. Enhancement of ABL kinase catalytic efficiency by a direct binding regulator is independent of other regulatory mechanisms. J. Biol. Chem. 2008, 283, 31401–31407. [Google Scholar] [CrossRef] [PubMed]
  204. Plattner, R.; Kadlec, L.; DeMali, K.A.; Kazlauskas, A.; Pendergast, A.M. c-Abl is activated by growth factors and Src family kinases and has a role in the cellular response to PDGF. Genes Dev. 1999, 13, 2400–2411. [Google Scholar] [CrossRef] [PubMed]
  205. Brasher, B.B.; Van Etten, R.A. c-Abl has high intrinsic tyrosine kinase activity that is stimulated by mutation of the Src homology 3 domain and by autophosphorylation at two distinct regulatory tyrosines. J. Biol. Chem. 2000, 275, 35631–35637. [Google Scholar] [CrossRef] [PubMed]
  206. Smith, K.M.; Van Etten, R.A. Activation of c-Abl kinase activity and transformation by a chemical inducer of dimerization. J. Biol. Chem. 2001, 276, 24372–24379. [Google Scholar] [CrossRef] [PubMed]
  207. Fan, P.D.; Cong, F.; Goff, S.P. Homo- and hetero-oligomerization of the c-Abl kinase and Abelson-interactor-1. Cancer Res. 2003, 63, 873–877. [Google Scholar] [PubMed]
  208. Sawyers, C.L.; McLaughlin, J.; Goga, A.; Havlik, M.; Witte, O. The nuclear tyrosine kinase c-Abl negatively regulates cell growth. Cell 1994, 77, 121–131. [Google Scholar] [CrossRef]
  209. Greuber, E.K.; Smith-Pearson, P.; Wang, J.; Pendergast, A.M. Role of ABL family kinases in cancer: From leukaemia to solid tumours. Nat. Rev. Cancer 2013, 13, 559–571. [Google Scholar] [CrossRef]
  210. Ganguly, S.S.; Fiore, L.S.; Sims, J.T.; Friend, J.W.; Srinivasan, D.; Thacker, M.A.; Cibull, M.L.; Wang, C.; Novak, M.; Kaetzel, D.M.; et al. c-Abl and Arg are activated in human primary melanomas, promote melanoma cell invasion via distinct pathways, and drive metastatic progression. Oncogene 2012, 31, 1804–1816. [Google Scholar] [CrossRef]
  211. Koos, B.; Jeibmann, A.; Lünenbürger, H.; Mertsch, S.; Nupponen, N.N.; Roselli, A.; Leuschner, I.; Paulus, W.; Frühwald, M.C.; Hasselblatt, M. The tyrosine kinase c-Abl promotes proliferation and is expressed in atypical teratoid and malignant rhabdoid tumors. Cancer 2010, 116, 5075–5081. [Google Scholar] [CrossRef] [PubMed]
  212. Ganguly, S.S.; Plattner, R. Activation of abl family kinases in solid tumors. Genes Cancer 2012, 3, 414–425. [Google Scholar] [CrossRef] [PubMed]
  213. Riccaboni, M.; Bianchi, I.; Petrillo, P. Spleen tyrosine kinases: Biology, therapeutic targets and drugs. Drug Discov. Today 2010, 15, 517–530. [Google Scholar] [CrossRef] [PubMed]
  214. Yanagi, S.; Inatome, R.; Takano, T.; Yamamura, H. Syk expression and novel function in a wide variety of tissues. Biochem. Biophys. Res. Commun. 2001, 288, 495–498. [Google Scholar] [CrossRef] [PubMed]
  215. Pugh, J.L.; Nemat-Gorgani, N.; Norman, P.J.; Guethlein, L.A.; Parham, P. Human NK Cells Downregulate Zap70 and Syk in Response to Prolonged Activation or DNA Damage. J. Immunol. 2018, 200, 1146–1158. [Google Scholar] [CrossRef] [PubMed]
  216. Paolini, R.; Molfetta, R.; Piccoli, M.; Frati, L.; Santoni, A. Ubiquitination and degradation of Syk and ZAP-70 protein tyrosine kinases in human NK cells upon CD16 engagement. Proc. Natl. Acad. Sci. USA 2001, 98, 9611–9616. [Google Scholar] [CrossRef] [PubMed]
  217. Woodside, D.G.; Obergfell, A.; Talapatra, A.; Calderwood, D.A.; Shattil, S.J.; Ginsberg, M.H. The N-terminal SH2 domains of Syk and ZAP-70 mediate phosphotyrosine-independent binding to integrin beta cytoplasmic domains. J. Biol. Chem. 2002, 277, 39401–39408. [Google Scholar] [CrossRef] [PubMed]
  218. Isakov, N.; Wange, R.L.; Burgess, W.H.; Watts, J.D.; Aebersold, R.; Samelson, L.E. ZAP-70 binding specificity to T cell receptor tyrosine-based activation motifs: The tandem SH2 domains of ZAP-70 bind distinct tyrosine-based activation motifs with varying affinity. J. Exp. Med. 1995, 181, 375–380. [Google Scholar] [CrossRef]
  219. Hobbs, H.T.; Shah, N.H.; Badroos, J.M.; Gee, C.L.; Marqusee, S.; Kuriyan, J. Differences in the dynamics of the tandem-SH2 modules of the Syk and ZAP-70 tyrosine kinases. Protein Sci. 2021, 30, 2373–2384. [Google Scholar] [CrossRef]
  220. Mócsai, A.; Ruland, J.; Tybulewicz, V.L. The SYK tyrosine kinase: A crucial player in diverse biological functions. Nat. Rev. Immunol. 2010, 10, 387–402. [Google Scholar] [CrossRef]
  221. Qu, C.; Zheng, D.; Li, S.; Liu, Y.; Lidofsky, A.; Holmes, J.A.; Chen, J.; He, L.; Wei, L.; Liao, Y.; et al. Tyrosine kinase SYK is a potential therapeutic target for liver fibrosis. Hepatology 2018, 68, 1125–1139. [Google Scholar] [CrossRef] [PubMed]
  222. Fitzer-Attas, C.J.; Schindler, D.G.; Waks, T.; Eshhar, Z. Harnessing Syk family tyrosine kinases as signaling domains for chimeric single chain of the variable domain receptors: Optimal design for T cell activation. J. Immunol. 1998, 160, 145–154. [Google Scholar] [CrossRef] [PubMed]
  223. Zhou, F.; Hu, J.; Ma, H.; Harrison, M.L.; Geahlen, R.L. Nucleocytoplasmic trafficking of the Syk protein tyrosine kinase. Mol. Cell. Biol. 2006, 26, 3478–3491. [Google Scholar] [CrossRef]
  224. Wange, R.L.; Malek, S.N.; Desiderio, S.; Samelson, L.E. Tandem SH2 domains of ZAP-70 bind to T cell antigen receptor zeta and CD3 epsilon from activated Jurkat T cells. J. Biol. Chem. 1993, 268, 19797–19801. [Google Scholar] [CrossRef] [PubMed]
  225. Bashore, F.M.; Katis, V.L.; Du, Y.; Sikdar, A.; Wang, D.; Bradshaw, W.J.; Rygiel, K.A.; Leisner, T.M.; Chalk, R.; Mishra, S.; et al. Characterization of covalent inhibitors that disrupt the interaction between the tandem SH2 domains of SYK and FCER1G phospho-ITAM. PLoS ONE 2024, 19, e0293548. [Google Scholar] [CrossRef] [PubMed]
  226. Fernández-Aguilar, L.M.; Vico-Barranco, I.; Arbulo-Echevarria, M.M.; Aguado, E. A Story of Kinases and Adaptors: The Role of Lck, ZAP-70 and LAT in Switch Panel Governing T-Cell Development and Activation. Biology 2023, 12, 1163. [Google Scholar] [CrossRef] [PubMed]
  227. Latour, S.; Chow, L.M.; Veillette, A. Differential intrinsic enzymatic activity of Syk and Zap-70 protein-tyrosine kinases. J. Biol. Chem. 1996, 271, 22782–22790. [Google Scholar] [CrossRef] [PubMed]
  228. Anbalagan, M.; Moroz, K.; Ali, A.; Carrier, L.; Glodowski, S.; Rowan, B.G. Subcellular localization of total and activated Src kinase in African American and Caucasian breast cancer. PLoS ONE 2012, 7, e33017. [Google Scholar] [CrossRef]
  229. Sudhesh Dev, S.; Zainal Abidin, S.A.; Farghadani, R.; Othman, I.; Naidu, R. Receptor Tyrosine Kinases and Their Signaling Pathways as Therapeutic Targets of Curcumin in Cancer. Front. Pharmacol. 2021, 12, 772510. [Google Scholar] [CrossRef]
  230. Wang, J.; Pendergast, A.M. The Emerging Role of ABL Kinases in Solid Tumors. Trends Cancer 2015, 1, 110–123. [Google Scholar] [CrossRef]
  231. Ren, R. Mechanisms of BCR-ABL in the pathogenesis of chronic myelogenous leukaemia. Nat. Rev. Cancer 2005, 5, 172–183. [Google Scholar] [CrossRef] [PubMed]
  232. Al Hamad, M. Contribution of BCR-ABL molecular variants and leukemic stem cells in response and resistance to tyrosine kinase inhibitors: A review. F1000Res 2021, 10, 1288. [Google Scholar] [CrossRef]
  233. Tan, X.; Yan, Y.; Song, B.; Zhu, S.; Mei, Q.; Wu, K. Focal adhesion kinase: From biological functions to therapeutic strategies. Exp. Hematol. Oncol. 2023, 12, 83. [Google Scholar] [CrossRef] [PubMed]
  234. Katoh, K. FAK-Dependent Cell Motility and Cell Elongation. Cells 2020, 9, 192. [Google Scholar] [CrossRef] [PubMed]
  235. Baillat, G.; Siret, C.; Delamarre, E.; Luis, J. Early adhesion induces interaction of FAK and Fyn in lipid domains and activates raft-dependent Akt signaling in SW480 colon cancer cells. Biochim. Biophys. Acta 2008, 1783, 2323–2331. [Google Scholar] [CrossRef] [PubMed]
  236. Oneyama, C.; Hikita, T.; Enya, K.; Dobenecker, M.W.; Saito, K.; Nada, S.; Tarakhovsky, A.; Okada, M. The lipid raft-anchored adaptor protein Cbp controls the oncogenic potential of c-Src. Mol. Cell 2008, 30, 426–436. [Google Scholar] [CrossRef] [PubMed]
  237. Garcia, R.; Bowman, T.L.; Niu, G.; Yu, H.; Minton, S.; Muro-Cacho, C.A.; Cox, C.E.; Falcone, R.; Fairclough, R.; Parsons, S.; et al. Constitutive activation of Stat3 by the Src and JAK tyrosine kinases participates in growth regulation of human breast carcinoma cells. Oncogene 2001, 20, 2499–2513. [Google Scholar] [CrossRef] [PubMed]
  238. Schwartzberg, P.L.; Finkelstein, L.D.; Readinger, J.A. TEC-family kinases: Regulators of T-helper-cell differentiation. Nat. Rev. Immunol. 2005, 5, 284–295. [Google Scholar] [CrossRef]
  239. Wang, F.; Hou, W.; Chitsike, L.; Xu, Y.; Bettler, C.; Perera, A.; Bank, T.; Cotler, S.J.; Dhanarajan, A.; Denning, M.F.; et al. ABL1, Overexpressed in Hepatocellular Carcinomas, Regulates Expression of NOTCH1 and Promotes Development of Liver Tumors in Mice. Gastroenterology 2020, 159, 289–305.e216. [Google Scholar] [CrossRef]
  240. Luttman, J.H.; Colemon, A.; Mayro, B.; Pendergast, A.M. Role of the ABL tyrosine kinases in the epithelial-mesenchymal transition and the metastatic cascade. Cell Commun. Signal. 2021, 19, 59. [Google Scholar] [CrossRef]
  241. Jain, A.; Tripathi, R.; Turpin, C.P.; Wang, C.; Plattner, R. Abl kinase regulation by BRAF/ERK and cooperation with Akt in melanoma. Oncogene 2017, 36, 4585–4596. [Google Scholar] [CrossRef] [PubMed]
  242. Singh, D.; Siddique, H.R. Epithelial-to-mesenchymal transition in cancer progression: Unraveling the immunosuppressive module driving therapy resistance. Cancer Metastasis Rev. 2024, 43, 155–173. [Google Scholar] [CrossRef] [PubMed]
  243. Ikeda, D.; Chi, S.; Uchiyama, S.; Nakamura, H.; Guo, Y.M.; Yamauchi, N.; Yuda, J.; Minami, Y. Molecular Classification and Overcoming Therapy Resistance for Acute Myeloid Leukemia with Adverse Genetic Factors. Int. J. Mol. Sci. 2022, 23, 5950. [Google Scholar] [CrossRef] [PubMed]
  244. Maruyama, I.N. Mechanisms of activation of receptor tyrosine kinases: Monomers or dimers. Cells 2014, 3, 304–330. [Google Scholar] [CrossRef] [PubMed]
  245. Trenker, R.; Jura, N. Receptor tyrosine kinase activation: From the ligand perspective. Curr. Opin. Cell Biol. 2020, 63, 174–185. [Google Scholar] [CrossRef] [PubMed]
  246. Raji, L.; Tetteh, A.; Amin, A.R.M.R. Role of c-Src in Carcinogenesis and Drug Resistance. Cancers 2023, 16, 32. [Google Scholar] [CrossRef]
  247. Penuel, E.; Martin, G.S. Transformation by v-Src: Ras-MAPK and PI3K-mTOR mediate parallel pathways. Mol. Biol. Cell 1999, 10, 1693–1703. [Google Scholar] [CrossRef] [PubMed]
  248. Lu, F.; Zhu, L.; Bromberger, T.; Yang, J.; Yang, Q.; Liu, J.; Plow, E.F.; Moser, M.; Qin, J. Mechanism of integrin activation by talin and its cooperation with kindlin. Nat. Commun. 2022, 13, 2362. [Google Scholar] [CrossRef]
  249. Li, Z.; Lee, H.; Zhu, C. Molecular mechanisms of mechanotransduction in integrin-mediated cell-matrix adhesion. Exp. Cell Res. 2016, 349, 85–94. [Google Scholar] [CrossRef]
  250. Luo, J.; Zou, H.; Guo, Y.; Tong, T.; Ye, L.; Zhu, C.; Deng, L.; Wang, B.; Pan, Y.; Li, P. SRC kinase-mediated signaling pathways and targeted therapies in breast cancer. Breast Cancer Res. 2022, 24, 99. [Google Scholar] [CrossRef]
  251. Kazemein Jasemi, N.S.; Ahmadian, M.R. Allosteric regulation of GRB2 modulates RAS activation. Small GTPases 2022, 13, 282–286. [Google Scholar] [CrossRef] [PubMed]
  252. Beadnell, T.C.; Nassar, K.W.; Rose, M.M.; Clark, E.G.; Danysh, B.P.; Hofmann, M.C.; Pozdeyev, N.; Schweppe, R.E. Src-mediated regulation of the PI3K pathway in advanced papillary and anaplastic thyroid cancer. Oncogenesis 2018, 7, 23. [Google Scholar] [CrossRef] [PubMed]
  253. Kanchanawong, P.; Calderwood, D.A. Organization, dynamics and mechanoregulation of integrin-mediated cell-ECM adhesions. Nat. Rev. Mol. Cell Biol. 2023, 24, 142–161. [Google Scholar] [CrossRef] [PubMed]
  254. Mitra, S.K.; Schlaepfer, D.D. Integrin-regulated FAK-Src signaling in normal and cancer cells. Curr. Opin. Cell Biol. 2006, 18, 516–523. [Google Scholar] [CrossRef] [PubMed]
  255. Li, S.; Sampson, C.; Liu, C.; Piao, H.L.; Liu, H.X. Integrin signaling in cancer: Bidirectional mechanisms and therapeutic opportunities. Cell Commun. Signal. 2023, 21, 266. [Google Scholar] [CrossRef] [PubMed]
  256. Yousefi, H.; Vatanmakanian, M.; Mahdiannasser, M.; Mashouri, L.; Alahari, N.V.; Monjezi, M.R.; Ilbeigi, S.; Alahari, S.K. Understanding the role of integrins in breast cancer invasion, metastasis, angiogenesis, and drug resistance. Oncogene 2021, 40, 1043–1063. [Google Scholar] [CrossRef] [PubMed]
  257. Huveneers, S.; Danen, E.H. Adhesion signaling-crosstalk between integrins, Src and Rho. J. Cell Sci. 2009, 122, 1059–1069. [Google Scholar] [CrossRef] [PubMed]
  258. Pang, X.; He, X.; Qiu, Z.; Zhang, H.; Xie, R.; Liu, Z.; Gu, Y.; Zhao, N.; Xiang, Q.; Cui, Y. Targeting integrin pathways: Mechanisms and advances in therapy. Signal Transduct. Target. Ther. 2023, 8, 1. [Google Scholar] [CrossRef] [PubMed]
  259. Playford, M.P.; Schaller, M.D. The interplay between Src and integrins in normal and tumor biology. Oncogene 2004, 23, 7928–7946. [Google Scholar] [CrossRef]
  260. Murphy, J.M.; Rodriguez, Y.A.R.; Jeong, K.; Ahn, E.E.; Lim, S.S. Targeting focal adhesion kinase in cancer cells and the tumor microenvironment. Exp. Mol. Med. 2020, 52, 877–886. [Google Scholar] [CrossRef]
  261. Zhou, J.; Yi, Q.; Tang, L. The roles of nuclear focal adhesion kinase (FAK) on Cancer: A focused review. J. Exp. Clin. Cancer Res. 2019, 38, 250. [Google Scholar] [CrossRef] [PubMed]
  262. Liu, J.; Zhang, Y.; Huang, H.; Lei, X.; Tang, G.; Cao, X.; Peng, J. Recent advances in Bcr-Abl tyrosine kinase inhibitors for overriding T315I mutation. Chem. Biol. Drug Des. 2021, 97, 649–664. [Google Scholar] [CrossRef] [PubMed]
  263. Brami-Cherrier, K.; Gervasi, N.; Arsenieva, D.; Walkiewicz, K.; Boutterin, M.C.; Ortega, A.; Leonard, P.G.; Seantier, B.; Gasmi, L.; Bouceba, T.; et al. FAK dimerization controls its kinase-dependent functions at focal adhesions. EMBO J. 2014, 33, 356–370. [Google Scholar] [CrossRef] [PubMed]
  264. Sulzmaier, F.J.; Jean, C.; Schlaepfer, D.D. FAK in cancer: Mechanistic findings and clinical applications. Nat. Rev. Cancer 2014, 14, 598–610. [Google Scholar] [CrossRef] [PubMed]
  265. Asai, A.; Miyata, Y.; Takehara, K.; Kanda, S.; Watanabe, S.I.; Greer, P.A.; Sakai, H. Pathological significance and prognostic significance of FES expression in bladder cancer vary according to tumor grade. J. Cancer Res. Clin. Oncol. 2018, 144, 21–31. [Google Scholar] [CrossRef] [PubMed]
  266. Zhang, S.; Chitu, V.; Stanley, E.R.; Elliott, B.E.; Greer, P.A. Fes tyrosine kinase expression in the tumor niche correlates with enhanced tumor growth, angiogenesis, circulating tumor cells, metastasis, and infiltrating macrophages. Cancer Res. 2011, 71, 1465–1473. [Google Scholar] [CrossRef] [PubMed]
  267. Delfino, F.J.; Stevenson, H.; Smithgall, T.E. A growth-suppressive function for the c-fes protein-tyrosine kinase in colorectal cancer. J. Biol. Chem. 2006, 281, 8829–8835. [Google Scholar] [CrossRef] [PubMed]
  268. Kanda, S.; Mochizuki, Y.; Miyata, Y.; Kanetake, H. The role of c-Fes in vascular endothelial growth factor-A-mediated signaling by endothelial cells. Biochem. Biophys. Res. Commun. 2003, 306, 1056–1063. [Google Scholar] [CrossRef]
  269. Miyata, Y.; Watanabe, S.; Matsuo, T.; Hayashi, T.; Sakai, H.; Xuan, J.W.; Greer, P.A.; Kanda, S. Pathological significance and predictive value for biochemical recurrence of c-Fes expression in prostate cancer. Prostate 2012, 72, 201–208. [Google Scholar] [CrossRef]
  270. Sangrar, W.; Zirgnibl, R.A.; Gao, Y.; Muller, W.J.; Jia, Z.; Greer, P.A. An identity crisis for fps/fes: Oncogene or tumor suppressor? Cancer Res. 2005, 65, 3518–3522. [Google Scholar] [CrossRef]
  271. Shen, L.; Zhang, C.; Cui, K.; Liang, X.; Zhu, G.; Hong, L. Fer-mediated activation of the Ras-MAPK signaling pathway drives the proliferation, migration, and invasion of endometrial carcinoma cells. Mol. Cell. Biochem. 2024, 479, 1787–1799. [Google Scholar] [CrossRef] [PubMed]
  272. Nir, U.; Grinshtain, E.; Breitbart, H. Fer and FerT: A New Regulatory Link between Sperm and Cancer Cells. Int. J. Mol. Sci. 2023, 24, 256. [Google Scholar] [CrossRef] [PubMed]
  273. Thomas, S.J.; Snowden, J.A.; Zeidler, M.P.; Danson, S.J. The role of JAK/STAT signalling in the pathogenesis, prognosis and treatment of solid tumours. Br. J. Cancer 2015, 113, 365–371. [Google Scholar] [CrossRef] [PubMed]
  274. Lopez-Bergami, P.; Fitchman, B.; Ronai, Z. Understanding signaling cascades in melanoma. Photochem. Photobiol. 2008, 84, 289–306. [Google Scholar] [CrossRef] [PubMed]
  275. Verhoeven, Y.; Tilborghs, S.; Jacobs, J.; De Waele, J.; Quatannens, D.; Deben, C.; Prenen, H.; Pauwels, P.; Trinh, X.B.; Wouters, A.; et al. The potential and controversy of targeting STAT family members in cancer. Semin. Cancer Biol. 2020, 60, 41–56. [Google Scholar] [CrossRef]
  276. Deng, C.; Liu, S.; Wang, Y.; Zhou, X.; Ma, H.; Wei, D.; Sun, S. The role of JAK/STAT signaling pathway and its inhibitors in diseases. Int. Immunopharmacol. 2020, 80, 106210. [Google Scholar] [CrossRef]
  277. Tucci, M.; Passarelli, A.; Mannavola, F.; Felici, C.; Stucci, L.S.; Cives, M.; Silvestris, F. Immune System Evasion as Hallmark of Melanoma Progression: The Role of Dendritic Cells. Front. Oncol. 2019, 9, 1148. [Google Scholar] [CrossRef] [PubMed]
  278. Logotheti, S.; Pützer, B.M. STAT3 and STAT5 Targeting for Simultaneous Management of Melanoma and Autoimmune Diseases. Cancers 2019, 11, 1448. [Google Scholar] [CrossRef] [PubMed]
  279. Lee, M.; Hirpara, J.L.; Eu, J.Q.; Sethi, G.; Wang, L.; Goh, B.C.; Wong, A.L. Targeting STAT3 and oxidative phosphorylation in oncogene-addicted tumors. Redox Biol. 2019, 25, 101073. [Google Scholar] [CrossRef]
  280. Sabaawy, H.E.; Ryan, B.M.; Khiabanian, H.; Pine, S.R. JAK/STAT of all trades: Linking inflammation with cancer development, tumor progression and therapy resistance. Carcinogenesis 2021, 42, 1411–1419. [Google Scholar] [CrossRef]
  281. Mullen, M.; Gonzalez-Perez, R.R. Leptin-Induced JAK/STAT Signaling and Cancer Growth. Vaccines 2016, 4, 26. [Google Scholar] [CrossRef] [PubMed]
  282. Khanna, P.; Chua, P.J.; Bay, B.H.; Baeg, G.H. The JAK/STAT signaling cascade in gastric carcinoma (Review). Int. J. Oncol. 2015, 47, 1617–1626. [Google Scholar] [CrossRef] [PubMed]
  283. Sugase, T.; Takahashi, T.; Serada, S.; Nakatsuka, R.; Fujimoto, M.; Ohkawara, T.; Hara, H.; Nishigaki, T.; Tanaka, K.; Miyazaki, Y.; et al. Suppressor of cytokine signaling-1 gene therapy induces potent antitumor effect in patient-derived esophageal squamous cell carcinoma xenograft mice. Int. J. Cancer 2017, 140, 2608–2621. [Google Scholar] [CrossRef] [PubMed]
  284. Zhao, G.; Liu, C.; Wen, X.; Luan, G.; Xie, L.; Guo, X. The translational values of TRIM family in pan-cancers: From functions and mechanisms to clinics. Pharmacol. Ther. 2021, 227, 107881. [Google Scholar] [CrossRef] [PubMed]
  285. Pencik, J.; Pham, H.T.; Schmoellerl, J.; Javaheri, T.; Schlederer, M.; Culig, Z.; Merkel, O.; Moriggl, R.; Grebien, F.; Kenner, L. JAK-STAT signaling in cancer: From cytokines to non-coding genome. Cytokine 2016, 87, 26–36. [Google Scholar] [CrossRef] [PubMed]
  286. Lu, Z.; Lu, C.; Li, C.; Jiao, Y.; Li, Y.; Zhang, G. Dracorhodin perchlorate induces apoptosis and G2/M cell cycle arrest in human esophageal squamous cell carcinoma through inhibition of the JAK2/STAT3 and AKT/FOXO3a pathways. Mol. Med. Rep. 2019, 20, 2091–2100. [Google Scholar] [CrossRef] [PubMed]
  287. Cho, S.C.; Lee, H.; Choi, B.Y. An updated review on molecular mechanisms underlying the anticancer effects of capsaicin. Food Sci. Biotechnol. 2017, 26, 1–13. [Google Scholar] [CrossRef] [PubMed]
  288. Liu, Y.; Wang, X.; Zeng, S.; Zhang, X.; Zhao, J.; Chen, X.; Yang, W.; Yang, Y.; Dong, Z.; Zhu, J.; et al. The natural polyphenol curcumin induces apoptosis by suppressing STAT3 signaling in esophageal squamous cell carcinoma. J. Exp. Clin. Cancer Res. 2018, 37, 303. [Google Scholar] [CrossRef] [PubMed]
  289. Hin Tang, J.J.; Hao Thng, D.K.; Lim, J.J.; Toh, T.B. JAK/STAT signaling in hepatocellular carcinoma. Hepat. Oncol. 2020, 7, HEP18. [Google Scholar] [CrossRef]
  290. Zhao, H.; Guo, Y.; Li, S.; Han, R.; Ying, J.; Zhu, H.; Wang, Y.; Yin, L.; Han, Y.; Sun, L.; et al. A novel anti-cancer agent Icaritin suppresses hepatocellular carcinoma initiation and malignant growth through the IL-6/Jak2/Stat3 pathway. Oncotarget 2015, 6, 31927–31943. [Google Scholar] [CrossRef]
  291. Ray, K.; Ujvari, B.; Ramana, V.; Donald, J. Cross-talk between EGFR and IL-6 drives oncogenic signaling and offers therapeutic opportunities in cancer. Cytokine Growth Factor Rev. 2018, 41, 18–27. [Google Scholar] [CrossRef] [PubMed]
  292. Morgan, E.L.; Macdonald, A. Manipulation of JAK/STAT Signalling by High-Risk HPVs: Potential Therapeutic Targets for HPV-Associated Malignancies. Viruses 2020, 12, 977. [Google Scholar] [CrossRef] [PubMed]
  293. Gkouveris, I.; Nikitakis, N.; Karanikou, M.; Rassidakis, G.; Sklavounou, A. JNK1/2 expression and modulation of STAT3 signaling in oral cancer. Oncol. Lett. 2016, 12, 699–706. [Google Scholar] [CrossRef] [PubMed]
  294. Bharadwaj, U.; Kasembeli, M.M.; Robinson, P.; Tweardy, D.J. Targeting Janus Kinases and Signal Transducer and Activator of Transcription 3 to Treat Inflammation, Fibrosis, and Cancer: Rationale, Progress, and Caution. Pharmacol. Rev. 2020, 72, 486–526. [Google Scholar] [CrossRef] [PubMed]
  295. O’Shea, J.J.; Schwartz, D.M.; Villarino, A.V.; Gadina, M.; McInnes, I.B.; Laurence, A. The JAK-STAT pathway: Impact on human disease and therapeutic intervention. Annu. Rev. Med. 2015, 66, 311–328. [Google Scholar] [CrossRef]
  296. Erdogan, F.; Radu, T.B.; Orlova, A.; Qadree, A.K.; de Araujo, E.D.; Israelian, J.; Valent, P.; Mustjoki, S.M.; Herling, M.; Moriggl, R.; et al. JAK-STAT core cancer pathway: An integrative cancer interactome analysis. J. Cell. Mol. Med. 2022, 26, 2049–2062. [Google Scholar] [CrossRef]
  297. Wang, S.; Xia, D.; Wang, X.; Cao, H.; Wu, C.; Sun, Z.; Zhang, D.; Liu, H. C/EBPβ regulates the JAK/STAT signaling pathway in triple-negative breast cancer. FEBS Open Bio 2021, 11, 1250–1258. [Google Scholar] [CrossRef]
  298. Kalimutho, M.; Parsons, K.; Mittal, D.; López, J.A.; Srihari, S.; Khanna, K.K. Targeted Therapies for Triple-Negative Breast Cancer: Combating a Stubborn Disease. Trends Pharmacol. Sci. 2015, 36, 822–846. [Google Scholar] [CrossRef]
  299. Rani, A.; Murphy, J.J. STAT5 in Cancer and Immunity. J. Interferon Cytokine Res. 2016, 36, 226–237. [Google Scholar] [CrossRef]
  300. Velloso, F.J.; Bianco, A.F.; Farias, J.O.; Torres, N.E.; Ferruzo, P.Y.; Anschau, V.; Jesus-Ferreira, H.C.; Chang, T.H.; Sogayar, M.C.; Zerbini, L.F.; et al. The crossroads of breast cancer progression: Insights into the modulation of major signaling pathways. OncoTargets Ther. 2017, 10, 5491–5524. [Google Scholar] [CrossRef]
  301. Yu, X.; Liu, J.; Qiu, H.; Hao, H.; Zhu, J.; Peng, S. Combined inhibition of ACK1 and AKT shows potential toward targeted therapy against KRAS-mutant non-small-cell lung cancer. Bosn. J. Basic Med. Sci. 2021, 21, 198–207. [Google Scholar] [CrossRef] [PubMed]
  302. Liu, F.; Wu, H. Identification of Prognostic Biomarkers and Molecular Targets Among JAK Family in Breast Cancer. J. Inflamm. Res. 2021, 14, 97–114. [Google Scholar] [CrossRef] [PubMed]
  303. Wang, B.; Song, K.; Chen, L.; Su, H.; Gao, L.; Liu, J.; Huang, A. Targeted inhibition of ACK1 can inhibit the proliferation of hepatocellular carcinoma cells through the PTEN/AKT/mTOR pathway. Cell Biochem. Funct. 2020, 38, 642–650. [Google Scholar] [CrossRef] [PubMed]
  304. Liu, X.; Wang, X.; Li, L.; Han, B. Research Progress of the Functional Role of ACK1 in Breast Cancer. Biomed. Res. Int. 2019, 2019, 1018034. [Google Scholar] [CrossRef]
  305. Peng, H.H.; Yang, H.C.; Rupa, D.; Yen, C.H.; Chiu, Y.W.; Yang, W.J.; Luo, F.J.; Yuan, T.C. ACK1 upregulated the proliferation of head and neck squamous cell carcinoma cells by promoting p27 phosphorylation and degradation. J. Cell Commun. Signal. 2022, 16, 567–578. [Google Scholar] [CrossRef] [PubMed]
  306. Lv, C.; Zhao, X.; Gu, H.; Huang, L.; Zhou, S.; Zhi, F. Involvement of Activated Cdc42 Kinase1 in Colitis and Colorectal Neoplasms. Med. Sci. Monit. 2016, 22, 4794–4802. [Google Scholar] [CrossRef] [PubMed]
  307. Kong, D.; Li, G.; Yang, Z.; Cheng, S.; Zhang, W.; Feng, L.; Zhang, K. Identification of an ACK1/TNK2-based prognostic signature for colon cancer to predict survival and inflammatory landscapes. BMC Cancer 2022, 22, 84. [Google Scholar] [CrossRef] [PubMed]
  308. Zhu, J.; Cao, K.; Zhao, M.; Ma, K.; Jiang, X.; Bai, Y.; Ling, X.; Ma, J. Improvement of ACK1-targeted therapy efficacy in lung adenocarcinoma using chloroquine or bafilomycin A1. Mol. Med. 2023, 29, 6. [Google Scholar] [CrossRef] [PubMed]
  309. Krisenko, M.O.; Geahlen, R.L. Calling in SYK: SYK’s dual role as a tumor promoter and tumor suppressor in cancer. Biochim. Biophys. Acta 2015, 1853, 254–263. [Google Scholar] [CrossRef]
  310. Torres-Hernandez, A.; Wang, W.; Nikiforov, Y.; Tejada, K.; Torres, L.; Kalabin, A.; Wu, Y.; Haq, M.I.U.; Khan, M.Y.; Zhao, Z.; et al. Targeting SYK signaling in myeloid cells protects against liver fibrosis and hepatocarcinogenesis. Oncogene 2019, 38, 4512–4526. [Google Scholar] [CrossRef]
  311. Wang, T.; Xu, Y.; Liu, X.; Zeng, Y.; Liu, L. miR-96-5p is the tumor suppressor in osteosarcoma via targeting SYK. Biochem. Biophys. Res. Commun. 2021, 572, 49–56. [Google Scholar] [CrossRef] [PubMed]
  312. Aguirre-Ducler, A.; Gianino, N.; Villarroel-Espindola, F.; Desai, S.; Tang, D.; Zhao, H.; Syrigos, K.; Trepicchio, W.L.; Kannan, K.; Gregory, R.C.; et al. Tumor cell SYK expression modulates the tumor immune microenvironment composition in human cancer via TNF-α dependent signaling. J. Immunother. Cancer 2022, 10, e005113. [Google Scholar] [CrossRef] [PubMed]
  313. Sung, Y.M.; Xu, X.; Sun, J.; Mueller, D.; Sentissi, K.; Johnson, P.; Urbach, E.; Seillier-Moiseiwitsch, F.; Johnson, M.D.; Mueller, S.C. Tumor suppressor function of Syk in human MCF10A in vitro and normal mouse mammary epithelium in vivo. PLoS ONE 2009, 4, e7445. [Google Scholar] [CrossRef]
  314. Rohila, D.; Park, I.H.; Pham, T.V.; Jones, R.; Tapia, E.; Liu, K.X.; Tamayo, P.; Yu, A.; Sharabi, A.B.; Joshi, S. Targeting macrophage Syk enhances responses to immune checkpoint blockade and radiotherapy in high-risk neuroblastoma. Front. Immunol. 2023, 14, 1148317. [Google Scholar] [CrossRef]
  315. Feng, G.; Wang, X. Role of spleen tyrosine kinase in the pathogenesis of chronic lymphocytic leukemia. Leuk. Lymphoma 2014, 55, 2699–2705. [Google Scholar] [CrossRef] [PubMed]
  316. Kassouf, T.; Larive, R.M.; Morel, A.; Urbach, S.; Bettache, N.; Marcial Medina, M.C.; Mèrezègue, F.; Freiss, G.; Peter, M.; Boissière-Michot, F.; et al. The Syk Kinase Promotes Mammary Epithelial Integrity and Inhibits Breast Cancer Invasion by Stabilizing the E-Cadherin/Catenin Complex. Cancers 2019, 11, 1974. [Google Scholar] [CrossRef]
  317. Abram, C.L.; Lowell, C.A. The expanding role for ITAM-based signaling pathways in immune cells. Sci. STKE 2007, 2007, re2. [Google Scholar] [CrossRef]
  318. Liu, D.; Mamorska-Dyga, A. Syk inhibitors in clinical development for hematological malignancies. J. Hematol. Oncol. 2017, 10, 145. [Google Scholar] [CrossRef] [PubMed]
  319. Singh, R.; Masuda, E.S.; Payan, D.G. Discovery and development of spleen tyrosine kinase (SYK) inhibitors. J. Med. Chem. 2012, 55, 3614–3643. [Google Scholar] [CrossRef]
  320. Thoma, G.; Veenstra, S.; Strang, R.; Blanz, J.; Vangrevelinghe, E.; Berghausen, J.; Lee, C.C.; Zerwes, H.G. Orally bioavailable Syk inhibitors with activity in a rat PK/PD model. Bioorg. Med. Chem. Lett. 2015, 25, 4642–4647. [Google Scholar] [CrossRef]
  321. Berg, L.J.; Finkelstein, L.D.; Lucas, J.A.; Schwartzberg, P.L. Tec family kinases in T lymphocyte development and function. Annu. Rev. Immunol. 2005, 23, 549–600. [Google Scholar] [CrossRef]
  322. Guo, W.; Liu, R.; Ono, Y.; Ma, A.H.; Martinez, A.; Sanchez, E.; Wang, Y.; Huang, W.; Mazloom, A.; Li, J.; et al. Molecular characteristics of CTA056, a novel interleukin-2-inducible T-cell kinase inhibitor that selectively targets malignant T cells and modulates oncomirs. Mol. Pharmacol. 2012, 82, 938–947. [Google Scholar] [CrossRef]
  323. Dawson, J.C.; Serrels, A.; Stupack, D.G.; Schlaepfer, D.D.; Frame, M.C. Targeting FAK in anticancer combination therapies. Nat. Rev. Cancer 2021, 21, 313–324. [Google Scholar] [CrossRef]
  324. Tang, K.J.; Constanzo, J.D.; Venkateswaran, N.; Melegari, M.; Ilcheva, M.; Morales, J.C.; Skoulidis, F.; Heymach, J.V.; Boothman, D.A.; Scaglioni, P.P. Focal Adhesion Kinase Regulates the DNA Damage Response and Its Inhibition Radiosensitizes Mutant KRAS Lung Cancer. Clin. Cancer Res. 2016, 22, 5851–5863. [Google Scholar] [CrossRef] [PubMed]
  325. Jamal, S.M.E.; Alamodi, A.; Wahl, R.U.; Grada, Z.; Shareef, M.A.; Hassan, S.Y.; Murad, F.; Hassan, S.L.; Santourlidis, S.; Gomez, C.R.; et al. Melanoma stem cell maintenance and chemo-resistance are mediated by CD133 signal to PI3K-dependent pathways. Oncogene 2020, 39, 5468–5478. [Google Scholar] [CrossRef] [PubMed]
  326. Hassan, M.; El Khattouti, A.; Ejaeidi, A.; Ma, T.; Day, W.A.; Espinoza, I.; Vijayakumar, S.; Gomez, C.R. Elevated Expression of Hepatoma Up-Regulated Protein Inhibits γ-Irradiation-Induced Apoptosis of Prostate Cancer Cells. J. Cell. Biochem. 2016, 117, 1308–1318. [Google Scholar] [CrossRef] [PubMed]
  327. Zhou, D.; Zuo, Y.; Pan, Z. Discovery of Potent and Highly Selective Interleukin-2-Inducible T-Cell Kinase Degraders with. J. Med. Chem. 2023, 66, 4979–4998. [Google Scholar] [CrossRef]
  328. Araujo, J.; Logothetis, C. Targeting Src signaling in metastatic bone disease. Int. J. Cancer 2009, 124, 1–6. [Google Scholar] [CrossRef]
  329. Belsches-Jablonski, A.P.; Biscardi, J.S.; Peavy, D.R.; Tice, D.A.; Romney, D.A.; Parsons, S.J. Src family kinases and HER2 interactions in human breast cancer cell growth and survival. Oncogene 2001, 20, 1465–1475. [Google Scholar] [CrossRef]
  330. Kim, L.C.; Song, L.; Haura, E.B. Src kinases as therapeutic targets for cancer. Nat. Rev. Clin. Oncol. 2009, 6, 587–595. [Google Scholar] [CrossRef]
  331. Susva, M.; Missbach, M.; Green, J. Src inhibitors: Drugs for the treatment of osteoporosis, cancer or both? Trends Pharmacol. Sci. 2000, 21, 489–495. [Google Scholar] [CrossRef]
  332. Missbach, M.; Altmann, E.; Susa, M. Tyrosine kinase inhibition in bone metabolism. Curr. Opin. Drug Discov. Dev. 2000, 3, 541–548. [Google Scholar]
  333. Hannon, R.A.; Clack, G.; Rimmer, M.; Swaisland, A.; Lockton, J.A.; Finkelman, R.D.; Eastell, R. Effects of the Src kinase inhibitor saracatinib (AZD0530) on bone turnover in healthy men: A randomized, double-blind, placebo-controlled, multiple-ascending-dose phase I trial. J. Bone Miner. Res. 2010, 25, 463–471. [Google Scholar] [CrossRef]
  334. Sundaramoorthi, R.; Kawahata, N.; Yang, M.G.; Shakespeare, W.C.; Metcalf, C.A.; Wang, Y.; Merry, T.; Eyermann, C.J.; Bohacek, R.S.; Narula, S.; et al. Structure-based design of novel nonpeptide inhibitors of the Src SH2 domain:phosphotyrosine mimetics exploiting multifunctional group replacement chemistry. Biopolymers 2003, 71, 717–729. [Google Scholar] [CrossRef] [PubMed]
  335. Hanke, J.H.; Gardner, J.P.; Dow, R.L.; Changelian, P.S.; Brissette, W.H.; Weringer, E.J.; Pollok, B.A.; Connelly, P.A. Discovery of a novel, potent, and Src family-selective tyrosine kinase inhibitor. Study of Lck- and FynT-dependent T cell activation. J. Biol. Chem. 1996, 271, 695–701. [Google Scholar] [CrossRef] [PubMed]
  336. Boyce, B.; Xing, L. Src inhibitors in the treatment of metastatic bone disease: Rationale and clinical data. Clin. Investig. 2011, 1, 1695–1706. [Google Scholar] [CrossRef]
  337. Zhu, S.; Wang, H.; Ranjan, K.; Zhang, D. Regulation, targets and functions of CSK. Front. Cell Dev. Biol. 2023, 11, 1206539. [Google Scholar] [CrossRef] [PubMed]
  338. Chüeh, A.C.; Advani, G.; Foroutan, M.; Smith, J.; Ng, N.; Nandurkar, H.; Lio, D.S.; Zhu, H.J.; Chong, Y.P.; Verkade, H.; et al. CSK-homologous kinase (CHK/MATK) is a potential colorectal cancer tumour suppressor gene epigenetically silenced by promoter methylation. Oncogene 2021, 40, 3015–3029. [Google Scholar] [CrossRef]
  339. Liu, X.; Du, Q.; Tian, C.; Tang, M.; Jiang, Y.; Wang, Y.; Cao, Y.; Wang, Z.; Yang, J.; Li, Y.; et al. Discovery of CAPE derivatives as dual EGFR and CSK inhibitors with anticancer activity in a murine model of hepatocellular carcinoma. Bioorg. Chem. 2021, 107, 104536. [Google Scholar] [CrossRef]
  340. Rossari, F.; Minutolo, F.; Orciuolo, E. Past, present, and future of Bcr-Abl inhibitors: From chemical development to clinical efficacy. J. Hematol. Oncol. 2018, 11, 84. [Google Scholar] [CrossRef]
  341. Giles, F.J.; O’Dwyer, M.; Swords, R. Class effects of tyrosine kinase inhibitors in the treatment of chronic myeloid leukemia. Leukemia 2009, 23, 1698–1707. [Google Scholar] [CrossRef]
  342. Weisberg, E.; Manley, P.; Mestan, J.; Cowan-Jacob, S.; Ray, A.; Griffin, J.D. AMN107 (nilotinib): A novel and selective inhibitor of BCR-ABL. Br. J. Cancer 2006, 94, 1765–1769. [Google Scholar] [CrossRef] [PubMed]
  343. Puttini, M.; Coluccia, A.M.; Boschelli, F.; Cleris, L.; Marchesi, E.; Donella-Deana, A.; Ahmed, S.; Redaelli, S.; Piazza, R.; Magistroni, V.; et al. In vitro and in vivo activity of SKI-606, a novel Src-Abl inhibitor, against imatinib-resistant Bcr-Abl+ neoplastic cells. Cancer Res. 2006, 66, 11314–11322. [Google Scholar] [CrossRef]
  344. Hochhaus, A.; Saussele, S.; Rosti, G.; Mahon, F.X.; Janssen, J.J.W.M.; Hjorth-Hansen, H.; Richter, J.; Buske, C.; Committee, E.G. Chronic myeloid leukaemia: ESMO Clinical Practice Guidelines for diagnosis, treatment and follow-up. Ann. Oncol. 2017, 28 (Suppl. S4), iv41–iv51. [Google Scholar] [CrossRef] [PubMed]
  345. Eskazan, A.E.; Keskin, D. Radotinib and its clinical potential in chronic-phase chronic myeloid leukemia patients: An update. Ther. Adv. Hematol. 2017, 8, 237–243. [Google Scholar] [CrossRef]
  346. Hughes, T.P.; Mauro, M.J.; Cortes, J.E.; Minami, H.; Rea, D.; DeAngelo, D.J.; Breccia, M.; Goh, Y.T.; Talpaz, M.; Hochhaus, A.; et al. Asciminib in Chronic Myeloid Leukemia after ABL Kinase Inhibitor Failure. N. Engl. J. Med. 2019, 381, 2315–2326. [Google Scholar] [CrossRef]
  347. Tan, F.H.; Putoczki, T.L.; Stylli, S.S.; Luwor, R.B. Ponatinib: A novel multi-tyrosine kinase inhibitor against human malignancies. OncoTargets Ther. 2019, 12, 635–645. [Google Scholar] [CrossRef] [PubMed]
  348. Gao, J.; Yao, Y.; Liu, C.; Xie, X.; Li, D.; Liu, P.; Wang, Z.; Zhang, B.; Ren, R. Synergism of FAK and ROS1 inhibitors in the treatment of. Int. J. Biol. Sci. 2023, 19, 2711–2724. [Google Scholar] [CrossRef]
  349. Amarante-Mendes, G.P.; Rana, A.; Datoguia, T.S.; Hamerschlak, N.; Brumatti, G. BCR-ABL1 Tyrosine Kinase Complex Signaling Transduction: Challenges to Overcome Resistance in Chronic Myeloid Leukemia. Pharmaceutics 2022, 14, 215. [Google Scholar] [CrossRef]
  350. Das, J.; Chen, P.; Norris, D.; Padmanabha, R.; Lin, J.; Moquin, R.V.; Shen, Z.; Cook, L.S.; Doweyko, A.M.; Pitt, S.; et al. 2-aminothiazole as a novel kinase inhibitor template. Structure-activity relationship studies toward the discovery of N-(2-chloro-6-methylphenyl)-2-[[6-[4-(2-hydroxyethyl)-1- piperazinyl)]-2-methyl-4-pyrimidinyl]amino)]-1,3-thiazole-5-carboxamide (dasatinib, BMS-354825) as a potent pan-Src kinase inhibitor. J. Med. Chem. 2006, 49, 6819–6832. [Google Scholar] [CrossRef]
  351. Li, H.; Zhang, P.; Liu, C.; Wang, Y.; Deng, Y.; Dong, W.; Yu, Y. The Structure, Function and Regulation of Protein Tyrosine Phosphatase Receptor Type J and Its Role in Diseases. Cells 2022, 12, 8. [Google Scholar] [CrossRef] [PubMed]
  352. Hale, A.J.; Ter Steege, E.; den Hertog, J. Recent advances in understanding the role of protein-tyrosine phosphatases in development and disease. Dev. Biol. 2017, 428, 283–292. [Google Scholar] [CrossRef] [PubMed]
  353. Xie, F.; Dong, H.; Zhang, H. Regulatory Functions of Protein Tyrosine Phosphatase Receptor Type O in Immune Cells. Front. Immunol. 2021, 12, 783370. [Google Scholar] [CrossRef] [PubMed]
  354. Bollu, L.R.; Mazumdar, A.; Savage, M.I.; Brown, P.H. Molecular Pathways: Targeting Protein Tyrosine Phosphatases in Cancer. Clin. Cancer Res. 2017, 23, 2136–2142. [Google Scholar] [CrossRef] [PubMed]
  355. Motiwala, T.; Jacob, S.T. Role of protein tyrosine phosphatases in cancer. Prog. Nucleic Acid Res. Mol. Biol. 2006, 81, 297–329. [Google Scholar] [CrossRef]
  356. Welsh, C.L.; Pandey, P.; Ahuja, L.G. Protein Tyrosine Phosphatases: A new paradigm in an old signaling system? Adv. Cancer Res. 2021, 152, 263–303. [Google Scholar] [CrossRef]
  357. Gao, P.P.; Qi, X.W.; Sun, N.; Sun, Y.Y.; Zhang, Y.; Tan, X.N.; Ding, J.; Han, F. The emerging roles of dual-specificity phosphatases and their specific characteristics in human cancer. Biochim. Biophys. Acta Rev. Cancer 2021, 1876, 188562. [Google Scholar] [CrossRef] [PubMed]
  358. Sivaganesh, V.; Scanlon, C.; Iskander, A.; Maher, S.; Lê, T.; Peethambaran, B. Protein Tyrosine Phosphatases: Mechanisms in Cancer. Int. J. Mol. Sci. 2021, 22, 12865. [Google Scholar] [CrossRef]
  359. Du, Y.; Grandis, J.R. Receptor-type protein tyrosine phosphatases in cancer. Chin. J. Cancer 2015, 34, 61–69. [Google Scholar] [CrossRef]
  360. Osman, N.; Lucas, S.; Cantrell, D. The role of tyrosine phosphorylation in the interaction of cellular tyrosine kinases with the T cell receptor zeta chain tyrosine-based activation motif. Eur. J. Immunol. 1995, 25, 2863–2869. [Google Scholar] [CrossRef]
  361. Stein-Gerlach, M.; Wallasch, C.; Ullrich, A. SHP-2, SH2-containing protein tyrosine phosphatase-2. Int. J. Biochem. Cell Biol. 1998, 30, 559–566. [Google Scholar] [CrossRef] [PubMed]
  362. Jadwin, J.A.; Curran, T.G.; Lafontaine, A.T.; White, F.M.; Mayer, B.J. Src homology 2 domains enhance tyrosine phosphorylation. J. Biol. Chem. 2018, 293, 623–637. [Google Scholar] [CrossRef] [PubMed]
  363. Song, Y.; Zhao, M.; Zhang, H.; Yu, B. Double-edged roles of protein tyrosine phosphatase SHP2 in cancer and its inhibitors in clinical trials. Pharmacol. Ther. 2022, 230, 107966. [Google Scholar] [CrossRef] [PubMed]
  364. Matozaki, T.; Murata, Y.; Saito, Y.; Okazawa, H.; Ohnishi, H. Protein tyrosine phosphatase SHP-2: A proto-oncogene product that promotes Ras activation. Cancer Sci. 2009, 100, 1786–1793. [Google Scholar] [CrossRef]
Figure 1. Structure of non-receptor tyrosine kinase families and their functional domains. Non-receptor tyrosine kinases (NRTKs) including Ack, Jak, Fes, Fak, Tec, Src, Csk, Abl, and Syk kinases. NRTKs consist of a single protein with N- and C-terminal regions. The N-terminus contains a kinase domain, which extends over approximately 300 residues, while the N-terminus of NRTKs is larger than the N-terminus. The structure and the functional domains and regions of different members of NRTKs include the following: Src homology (SH) domains that are referred to as SH4, SH3, SH2, and catalytic SH1 domains, the Pleckstrin homology (PH), four-point-one, ezrin, radixin, moesin (FERM), the Janus homology 2 (JH2) domain, and the Fes/Fer/Cdc-42-interacting protein homology (FCH) domains, Bruton’s tyrosine kinase (Btk)-like zinc finger, the coiled-coil motifs (CC), and proline-rich region (pr). The Cdc42/Rac-interactive (CRIB) domains, DNA-binding domains (DNA), action-binding domain, and a focal adhesion targeting domain (FAT).
Figure 1. Structure of non-receptor tyrosine kinase families and their functional domains. Non-receptor tyrosine kinases (NRTKs) including Ack, Jak, Fes, Fak, Tec, Src, Csk, Abl, and Syk kinases. NRTKs consist of a single protein with N- and C-terminal regions. The N-terminus contains a kinase domain, which extends over approximately 300 residues, while the N-terminus of NRTKs is larger than the N-terminus. The structure and the functional domains and regions of different members of NRTKs include the following: Src homology (SH) domains that are referred to as SH4, SH3, SH2, and catalytic SH1 domains, the Pleckstrin homology (PH), four-point-one, ezrin, radixin, moesin (FERM), the Janus homology 2 (JH2) domain, and the Fes/Fer/Cdc-42-interacting protein homology (FCH) domains, Bruton’s tyrosine kinase (Btk)-like zinc finger, the coiled-coil motifs (CC), and proline-rich region (pr). The Cdc42/Rac-interactive (CRIB) domains, DNA-binding domains (DNA), action-binding domain, and a focal adhesion targeting domain (FAT).
Cancers 16 02754 g001
Figure 2. An overview of the domain organization of Ack1. The different structural domains of Ack1 are shown. These domains include the sterile alpha motif (SAM); the nuclear transport signal (NES); catalytic kinase domain (Cat. K. domain); Src homology domain 3 (SH3); Cdc42; Rac-interactive binding domain (CRIB); Cdc42 binding domain (CB); CL, clathrin-binding domain (CB), proline-rich region, Mig6 homology region (MHR); UBA, ubiquitin association domain (UBA).
Figure 2. An overview of the domain organization of Ack1. The different structural domains of Ack1 are shown. These domains include the sterile alpha motif (SAM); the nuclear transport signal (NES); catalytic kinase domain (Cat. K. domain); Src homology domain 3 (SH3); Cdc42; Rac-interactive binding domain (CRIB); Cdc42 binding domain (CB); CL, clathrin-binding domain (CB), proline-rich region, Mig6 homology region (MHR); UBA, ubiquitin association domain (UBA).
Cancers 16 02754 g002
Figure 3. Structure and functional domains of JAKs. Each member of the JAKs (JAK1, JAK2, JAK3, and TYK2) consists of seven homology domains (JH) organized into four different domains, including the kinase domain that constitutes the JH1, the pseudo domain that constitutes JH2, the Src homology2 (SH2) that constitutes both JH3 and JH4, and the four-point-one, ezrin, radixin, and moesin (FERM) domain that constitutes JH5, JH5, JH6, and JH7. The phosphorylation of JAK1 occurs at Tyr1038/Tyr1039 residues; the phosphorylation of JAK2 occurs at -Tyr1007/Tyr1008; the phosphorylation of JAK3 occurs at Tyr980/Tyr981 residues; and the phosphorylation of Tyk2 occurs at Tyr1054/Tyr1055 residues.
Figure 3. Structure and functional domains of JAKs. Each member of the JAKs (JAK1, JAK2, JAK3, and TYK2) consists of seven homology domains (JH) organized into four different domains, including the kinase domain that constitutes the JH1, the pseudo domain that constitutes JH2, the Src homology2 (SH2) that constitutes both JH3 and JH4, and the four-point-one, ezrin, radixin, and moesin (FERM) domain that constitutes JH5, JH5, JH6, and JH7. The phosphorylation of JAK1 occurs at Tyr1038/Tyr1039 residues; the phosphorylation of JAK2 occurs at -Tyr1007/Tyr1008; the phosphorylation of JAK3 occurs at Tyr980/Tyr981 residues; and the phosphorylation of Tyk2 occurs at Tyr1054/Tyr1055 residues.
Cancers 16 02754 g003
Figure 4. Structure and functional domains of FES/FER kinases. FES and FER share a similar domain organization, including F-Bin–Amphiphysin–Rvs (F-BAR), F-BAR extension (FX), Src homology2 (SH2), and protein tyrosine kinase (PTK) domains. Fes is structured as a dimer mediated by its F-BAR domain that can exist in higher-order oligomers. The N-terminal domain of Fes is to repress its kinase activity, which can be attenuated by proline insertion mutations, leading to the destruction of oligomerization. As a consequence, the repressed Fes oligomer undergoes significant conformation via the interaction of the F-BAR domain with the phosphatidylinositol 4,5-bisphosphate or FX domain binding to phosphatidic acid, which can be produced by phospholipase D. Accordingly, Fes more readily interacts with SH2 ligands, which leads to kinase activation via the SH2–PTK interface.
Figure 4. Structure and functional domains of FES/FER kinases. FES and FER share a similar domain organization, including F-Bin–Amphiphysin–Rvs (F-BAR), F-BAR extension (FX), Src homology2 (SH2), and protein tyrosine kinase (PTK) domains. Fes is structured as a dimer mediated by its F-BAR domain that can exist in higher-order oligomers. The N-terminal domain of Fes is to repress its kinase activity, which can be attenuated by proline insertion mutations, leading to the destruction of oligomerization. As a consequence, the repressed Fes oligomer undergoes significant conformation via the interaction of the F-BAR domain with the phosphatidylinositol 4,5-bisphosphate or FX domain binding to phosphatidic acid, which can be produced by phospholipase D. Accordingly, Fes more readily interacts with SH2 ligands, which leads to kinase activation via the SH2–PTK interface.
Cancers 16 02754 g004
Figure 5. (A) Structure of focal adhesion kinase (FAK) and its homologous FAK-related proline-rich tyrosine kinase 2 (Pyk2). Both FAK and PyK2 kinases consists of three domains, including the four-point-one, ezrin, radixin, and moesin (FERM) that is divided into three subdomains, referred to as F1, F2, and F3, central kinase domain, and focal adhesion targeting (FAT) domain. Both FAK and Pyk2 kinases contain a nuclear localization sequence (NLS) and a nuclear export sequence (NES) in addition to several shared conserved phosphorylation sites of tyrosine (Tyr) residues that are in the FAK protein sequence at Tyr397, Tyr576, and Tyr577 residues, while in PyK2, the phosphorylation sites of tyrosine are located at Tyr402, Tyr579, and Tyr580 residues. Both FAK and PyK2 kinases possess a lysine (Lys) mutation in the kinase domain of FAK at Lys454 residue and in the kinase domain of Pyk2 at Lys457 residue. (B) Src/FAK-mediated transduction pathways contribute to cancer progression. FAK is activated by integrins to enhance the activity of phosphoinositide 3-kinase (PI3K). The activation of FAK-dependent pathways is associated with the stimulation of cell proliferation, angiogenesis, and survival. Upon the phosphorylation of FAK, we are able to recruit Grb2 and the p85 regulatory subunit of PI3K to enhance the activation of both the Grb2/Ras/MAPK and PI3K/Akt pathways, respectively, to promote tumor angiogenesis, survival, and proliferation.
Figure 5. (A) Structure of focal adhesion kinase (FAK) and its homologous FAK-related proline-rich tyrosine kinase 2 (Pyk2). Both FAK and PyK2 kinases consists of three domains, including the four-point-one, ezrin, radixin, and moesin (FERM) that is divided into three subdomains, referred to as F1, F2, and F3, central kinase domain, and focal adhesion targeting (FAT) domain. Both FAK and Pyk2 kinases contain a nuclear localization sequence (NLS) and a nuclear export sequence (NES) in addition to several shared conserved phosphorylation sites of tyrosine (Tyr) residues that are in the FAK protein sequence at Tyr397, Tyr576, and Tyr577 residues, while in PyK2, the phosphorylation sites of tyrosine are located at Tyr402, Tyr579, and Tyr580 residues. Both FAK and PyK2 kinases possess a lysine (Lys) mutation in the kinase domain of FAK at Lys454 residue and in the kinase domain of Pyk2 at Lys457 residue. (B) Src/FAK-mediated transduction pathways contribute to cancer progression. FAK is activated by integrins to enhance the activity of phosphoinositide 3-kinase (PI3K). The activation of FAK-dependent pathways is associated with the stimulation of cell proliferation, angiogenesis, and survival. Upon the phosphorylation of FAK, we are able to recruit Grb2 and the p85 regulatory subunit of PI3K to enhance the activation of both the Grb2/Ras/MAPK and PI3K/Akt pathways, respectively, to promote tumor angiogenesis, survival, and proliferation.
Cancers 16 02754 g005
Figure 6. Structure of TEC family kinases. The five kinases of the TEC family are represented by a diagram showing their structural domains. PH: Pleckstrin homology domain; TH: Tec homology domain; BH: Btk homology motif PR: proline-rich region; Cys: cysteine-rich sequence; N-Myr: N-terminal myristylation signal; SH: Src homology domain (SH3, SH2, and kinase) domains.
Figure 6. Structure of TEC family kinases. The five kinases of the TEC family are represented by a diagram showing their structural domains. PH: Pleckstrin homology domain; TH: Tec homology domain; BH: Btk homology motif PR: proline-rich region; Cys: cysteine-rich sequence; N-Myr: N-terminal myristylation signal; SH: Src homology domain (SH3, SH2, and kinase) domains.
Cancers 16 02754 g006
Figure 7. Structure of SRC family kinases, including their functional domains and dependent signals. Activation of Src family kinases is a membrane-dependent mechanism and is regulated by lipid/myristate modification within the SH4 domain and membrane binding. The modification of SH4 by myristylation and/or palmytoilation is to facilitate the localization of Src family kinases to the cell membrane, and the phosphorylation of tyrosine residues within the kinase domain to enhance Src activity. The inhibition of Src family kinases is mediated via ubiquitination sites, lysine 429 (Lys429) residue, and/or by the phosphorylation of tyrosine (Tyr) residues located in C-regulatory tail. Domain/regulatory regions are shown as boxes: Src homology 1 (SH1); tyrosine kinase/catalytic domain; Src homology 2 (SH2); Src homology 3 (SH3); Src homology 4 (SH4); and unstructured unique domain (UD).
Figure 7. Structure of SRC family kinases, including their functional domains and dependent signals. Activation of Src family kinases is a membrane-dependent mechanism and is regulated by lipid/myristate modification within the SH4 domain and membrane binding. The modification of SH4 by myristylation and/or palmytoilation is to facilitate the localization of Src family kinases to the cell membrane, and the phosphorylation of tyrosine residues within the kinase domain to enhance Src activity. The inhibition of Src family kinases is mediated via ubiquitination sites, lysine 429 (Lys429) residue, and/or by the phosphorylation of tyrosine (Tyr) residues located in C-regulatory tail. Domain/regulatory regions are shown as boxes: Src homology 1 (SH1); tyrosine kinase/catalytic domain; Src homology 2 (SH2); Src homology 3 (SH3); Src homology 4 (SH4); and unstructured unique domain (UD).
Cancers 16 02754 g007
Figure 8. Structure and domains of human SRMS, BRK, FRK, c-Src, Csk, IYK/BSK, and chicken v-Src. These kinases constitute SH3, SH2, and the kinase domain on their N-terminal region. Key residues, which are essential for the regulation of the enzymatic activity of these kinases, are highlighted. These include SRMS Tyr380 of the SRMS, Tyr342 of BRK, Tyr387 of FRK, Tyr419 of c-Src, Tyr497 of IYK/BSK, and Tyr416 of v-Src. The C-terminal regulatory tyrosine residues include Tyr447 of BRK, Tyr497 of FRK, Tyr530 of c-Src, and Tyr504 of IYK/BSK. The ATP-contacting lysine residues include Lys258 of SRMS, Lys219 of BRK, Lys262 of FRK, Lys298 of c-Src, Lys222 of Csk, Lys269 of IYK/BSK, and Lys295 of v-Src Lys295. All numbering is for human proteins, and v-Src is from a retroviral source with its own numbering.
Figure 8. Structure and domains of human SRMS, BRK, FRK, c-Src, Csk, IYK/BSK, and chicken v-Src. These kinases constitute SH3, SH2, and the kinase domain on their N-terminal region. Key residues, which are essential for the regulation of the enzymatic activity of these kinases, are highlighted. These include SRMS Tyr380 of the SRMS, Tyr342 of BRK, Tyr387 of FRK, Tyr419 of c-Src, Tyr497 of IYK/BSK, and Tyr416 of v-Src. The C-terminal regulatory tyrosine residues include Tyr447 of BRK, Tyr497 of FRK, Tyr530 of c-Src, and Tyr504 of IYK/BSK. The ATP-contacting lysine residues include Lys258 of SRMS, Lys219 of BRK, Lys262 of FRK, Lys298 of c-Src, Lys222 of Csk, Lys269 of IYK/BSK, and Lys295 of v-Src Lys295. All numbering is for human proteins, and v-Src is from a retroviral source with its own numbering.
Cancers 16 02754 g008
Figure 9. (A) Structural and functional domains of the Abl kinases. Alternative splicing of both Abl1 and Abl2 results in the production of protein variants/isoforms. The isoforms of Abl1 and Abl2, namely Abl1a and Abl2b, are characterized by their myristylation at the N-termini. Abl1 possesses two splice variants, Abl1a and Abl1b, and Abl2 also possesses two splice variants, Abl2a and Abl2b. The N-terminal region of Abl constitutes SH3, SH2, and SH1/kinase domains. The C- terminus of Abl constitutes conserved filamentous (F)-actin-binding (BD), globular (G)-actin-binding, and DNA-binding domains, and three nuclear localization signal (NLS) motifs and one nuclear export signal (NES). Both Abl1 and Abl2 kinases share the same domains in their N-terminus and are different in terms of their C-terminus. The C-terminus of Abl2 constitutes two BD domains and a microtubule (MT)-binding domain. Both Abl1 and Abl2 have conserved PXXP motifs to facilitate protein–protein interactions and tyrosine phosphorylation sites as indicated. (B) Three-dimensional conformation of Abl kinase protein. The different domains of the ABL kinase including SH3 (Green), SH2 (red), and SH1/kinase (yellow) domains, and the catalytic pocket, allosteric pocket, activation loop, and the carboxy-terminal helix are demonstrated.
Figure 9. (A) Structural and functional domains of the Abl kinases. Alternative splicing of both Abl1 and Abl2 results in the production of protein variants/isoforms. The isoforms of Abl1 and Abl2, namely Abl1a and Abl2b, are characterized by their myristylation at the N-termini. Abl1 possesses two splice variants, Abl1a and Abl1b, and Abl2 also possesses two splice variants, Abl2a and Abl2b. The N-terminal region of Abl constitutes SH3, SH2, and SH1/kinase domains. The C- terminus of Abl constitutes conserved filamentous (F)-actin-binding (BD), globular (G)-actin-binding, and DNA-binding domains, and three nuclear localization signal (NLS) motifs and one nuclear export signal (NES). Both Abl1 and Abl2 kinases share the same domains in their N-terminus and are different in terms of their C-terminus. The C-terminus of Abl2 constitutes two BD domains and a microtubule (MT)-binding domain. Both Abl1 and Abl2 have conserved PXXP motifs to facilitate protein–protein interactions and tyrosine phosphorylation sites as indicated. (B) Three-dimensional conformation of Abl kinase protein. The different domains of the ABL kinase including SH3 (Green), SH2 (red), and SH1/kinase (yellow) domains, and the catalytic pocket, allosteric pocket, activation loop, and the carboxy-terminal helix are demonstrated.
Cancers 16 02754 g009
Figure 10. Structure of spleen tyrosine kinase (Syk)/Syk-B and ZAP-70 family protein tyrosine kinases. Syk kinases is organized in three functional domains, including two Src-homology 2 (SH2) domains and kinase domains. The three functional domains of the Syk kinase are connected via two A and B intermediate domains. Tyrosine residues of Syk that undergo phosphorylation are indicated. These tyrosine residues are essential for the regulation of enzymatic activity and to recruit other signaling proteins. The phosphorylation of Tyr342 and Tyr346 is dependent on the Src family kinases (SFK); phosphorylation of Tyr130, Tyr317, Tyr519, and Tyr520 is dependent on Syk itself and is target for autophosphorylation.
Figure 10. Structure of spleen tyrosine kinase (Syk)/Syk-B and ZAP-70 family protein tyrosine kinases. Syk kinases is organized in three functional domains, including two Src-homology 2 (SH2) domains and kinase domains. The three functional domains of the Syk kinase are connected via two A and B intermediate domains. Tyrosine residues of Syk that undergo phosphorylation are indicated. These tyrosine residues are essential for the regulation of enzymatic activity and to recruit other signaling proteins. The phosphorylation of Tyr342 and Tyr346 is dependent on the Src family kinases (SFK); phosphorylation of Tyr130, Tyr317, Tyr519, and Tyr520 is dependent on Syk itself and is target for autophosphorylation.
Cancers 16 02754 g010
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Eshaq, A.M.; Flanagan, T.W.; Hassan, S.-Y.; Al Asheikh, S.A.; Al-Amoudi, W.A.; Santourlidis, S.; Hassan, S.-L.; Alamodi, M.O.; Bendhack, M.L.; Alamodi, M.O.; et al. Non-Receptor Tyrosine Kinases: Their Structure and Mechanistic Role in Tumor Progression and Resistance. Cancers 2024, 16, 2754. https://doi.org/10.3390/cancers16152754

AMA Style

Eshaq AM, Flanagan TW, Hassan S-Y, Al Asheikh SA, Al-Amoudi WA, Santourlidis S, Hassan S-L, Alamodi MO, Bendhack ML, Alamodi MO, et al. Non-Receptor Tyrosine Kinases: Their Structure and Mechanistic Role in Tumor Progression and Resistance. Cancers. 2024; 16(15):2754. https://doi.org/10.3390/cancers16152754

Chicago/Turabian Style

Eshaq, Abdulaziz M., Thomas W. Flanagan, Sofie-Yasmin Hassan, Sara A. Al Asheikh, Waleed A. Al-Amoudi, Simeon Santourlidis, Sarah-Lilly Hassan, Maryam O. Alamodi, Marcelo L. Bendhack, Mohammed O. Alamodi, and et al. 2024. "Non-Receptor Tyrosine Kinases: Their Structure and Mechanistic Role in Tumor Progression and Resistance" Cancers 16, no. 15: 2754. https://doi.org/10.3390/cancers16152754

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop